Você está na página 1de 157

E d u c a tio n a n d C u ltu r e

Pilot Project No: CZ/02/B/F/PP/134001

Course on Dynamics of Multidisplicinary Systems

Part I

c
2008

Herman Mann
Czech Technical University in Prague
herman.j.mann@gmail.com
Contents

1 Introduction to dynamics 1-1


1.1 Modeling and simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-1
1.2 Role of simulation in engineering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-2

2 Rectilinear mechanical systems 2-1


2.1 Rectilinear systems and their surroundings . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-1
2.2 Rectilinear interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-2
2.3 Composition of rectilinear interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-4
2.4 Rectilinear inertia of mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-6
2.5 Prescribed rectilinear interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-8
2.6 Rectilinear damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-12
2.7 Rectilinear compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-14
2.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-16

3 Rotational mechanical systems 3-1


3.1 Rotational interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-1
3.2 Ideal rotational inertors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-2
3.3 Ideal rotational dampers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-2
3.4 Ideal rotational springs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-4

4 Coupled mechanical systems 4-1


4.1 Translatory-to-translatory couplings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-1
4.2 Rotary-to-rotary couplings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-3
4.3 Translatory-to-rotary couplings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-8
4.4 Load conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-11

5 Planar mechanical systems 5-1


5.1 Planar motion of particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-1
5.2 Planar systems with massless rods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-7

6 Electrical systems 6-1


6.1 Electrical interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-1
6.2 Prescribed electrical interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-2
6.3 Pure electrical resistors and conductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-3
6.4 Electrical capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-4
6.5 Pure electrical inductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-5
6.6 Coupled inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-7
6.7 More coupled inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-8
6.8 Coupled inductors and ideal transformer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-9

7 Fluid systems 7-1


7.1 Fluid energy interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-1
7.2 Modeling of fluid systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-3
7.3 Fluid accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-5
7.4 Fluid flow restrictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-9
7.5 Fluid inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-12
7.6 Fluid transducers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-13

1
2 CONTENTS

8 Thermal systems 8-1


8.1 Thermal interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-1
8.2 Composition of thermal interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-2
8.3 Accumulation of heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-2
8.4 Heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-3
8.5 Thermal inductors and transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-5
8.6 Thermoelectric transducers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-5

9 Magnetic systems 9-1


9.1 Magnetic interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-1
9.2 Pure magnetic capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-2
9.3 Pure magnetic conductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-4
9.4 Ideal magnetic inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-4
9.5 Magneto-electric conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-4
9.6 Magnetic nonlinearities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-6
9.7 Permanent magnet excited systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-7

10 Electromechanical systems 10-1


10.1 Conductor moving in magnetic field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-1
10.2 Coil moving in magnetic field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-1
10.3 Single-phase alternator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-2
10.4 Permanent-magnet dc machine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-3
10.5 Movable-plate condenser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-4
10.6 Movable-core solenoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-5
10.7 Electromagnetic relay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-6
10.8 Synchronous reluctance machine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-7
10.9 Variable-reluctance stepping motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-7

11 Unified approach to modeling 11-1


11.1 Dynamic behavior of systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-1
11.2 Dissipative and accumulative elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-5
11.3 Sources of energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-7
11.4 Orientation of physical elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-10
11.5 Configurations of physical elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-12
11.6 Energy transducers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-14

12 Formulation of system equations 12-1


12.1 Full set of equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-1
12.2 Reduced equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-2
12.3 Extended nodal formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-4
12.4 Block diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-8
12.5 Lagrange’s equations of multipoles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-11

13 Formulation of transfer functions 13-1


13.1 Formulation of system-response transforms . . . . . . . . . . . . . . . . . . . . . . . . . . 13-1
13.2 Determinant evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13-1
13.3 Formulation of system transfer functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13-2
13.4 Solved examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13-2
About this course

All aspects of modern technology are underlined by the subject of system dynamics, which plays the
determining role in the market competition of engineering products. Its importance increases with the
ever-growing demands on operational speed, energetic efficiency, safety, reliability, and environment pro-
tection. At the same time, the engineering products become more and more complex with regards both
to the number of their components as well as to the variety of phenomena affecting the product dynamics,
either in a useful or undesirable way.
To understand and predict dynamic behavior of existing systems as well as to design new systems,
engineers resort to computer-assisted modeling, simulation and analysis of the systems. Modeling is
in this context a procedure for forming a simplified representation of those features of real dynamic
systems that are relevant to their dynamic behavior. The simplification is based on an abstraction of
the relevant system features and on their idealization. The resulting model representation can take the
form of equations, diagrams, or tables of data, for example. Simulation is a procedure imitating dynamic
behavior of the modelled real systems. Analysis is a procedure investigating changes of the system
behavior if the system or its surroundings are changing so it gives more thorough information about the
system dynamics than simulation.
Such techniques can replace to a certain degree expensive and time consuming design verification
based on construction of system prototypes and their experimental investigation using measuring instru-
ments. At the same time, these techniques allow engineers considering a larger variety of system design
alternatives and their more thorough verification. Design verification by simulation is the only option in
cases where the experimentation with real systems would be too expensive or too dangerous. Analysis
can also simplify the system operation and maintenance by detecting an impending system failure as well
as by helping to locate a fault in the system and to diagnose its cause. As these computation techniques
allow for faster introduction of a new product into the market as well as for attaining and maintaining
the product higher quality, they help in acquiring higher profits.
This Web-based course on dynamics of multidisciplinary controlled systems has been developed for

• regular students wishing to complement the traditional face-to-face courses

• distance-education students at different levels of vocational study and training

• practising engineers in the context of continuing education or lifelong learning

• teachers intending to innovate the courses they teach

• industrial enterprisers interested in enhancing the qualification and efficiency of their personnel

• providers of regular, continuing and life-long-learning courses

The subject of dynamics and control underlies all aspects of modern technology and plays the de-
termining role in the world-market competition of engineering products. Its importance increases with
the ever-growing demands on operational speed, efficiency, safety, reliability, or environment protection.
Nevertheless, entrepreneurs report lack of professionals sufficiently qualified in this field and vocational
schools complain about the overall decline of interest in engineering study among young people. Profes-
sional associations call for radical changes in the engineering curriculum and for innovative approaches
to vocational training.
The existing courses on dynamics and control are criticized for discouraging young people from en-
gineering study by overemphasizing theory and mathematics at the expense of practical engineering
issues. Dynamics is usually covered in several courses separated along the borders between the tradi-
tional engineering disciplines despite the fact that most of the contemporary engineering products are of

3
4 CONTENTS

multidisciplinary nature. Control courses are criticized for presenting ’textbook’ problems engineered to
fit the theory without undertaking a realistic modeling of the controlled systems. Computers are often
used to carry out old exercises without radical modification of the curriculum to fully incorporate current
computer capabilities.
The contents, emphasis and style of this course differs from most of the existing courses by a number of
factors. The first part of the course exposes learners in a systematic and unified way to realistic modelling
and simulation of multidisciplinary systems. This approach is supported by a free online access to the
simulation software DYNAST. It allows for submitting dynamic models directly in a graphical form
isomorphic to the geometric configuration of the modelled real systems. The underlying equations are
not only solved, but also automatically formulated.
Using this tool, learners can be introduced to dynamics through modeling and simulation of simple
yet practical examples. This stimulates their interest before they are exposed to more abstract subjects
like equation formulation, block-diagram construction, Laplace transformation, etc. In addition, learners
are given in the course a better ‘feel’ for system dynamics by graphical visualization, 3D animation and
interactive virtual experiments. Computers are thus integrated into the course curriculum from the very
start as learning tools without distracting learners from dynamics by issues not related directly to it.
In the same way, learners are given the hands-on opportunity to acquire the skills necessary for efficient
solving real-life problems in their engineering profession.
Different target groups can select individual paths through the course, paths tailor-made to their
previous experience and respecting their learning needs. Depending on their background, learners can
choose to start from the module on dynamics of either mechanical, electrical or fluid systems. More
advanced learners can go directly to the dynamics of multidisciplinary systems, or even to control. The
course material as well as the learning environment suit both to self-study and remote tutoring. They
both strongly support investigative and collaborative modes of learning.
Learners are given the opportunity to benefit from ‘organizational learning’, i.e. from utilizing know-
ledge recorded during previous problem solving both in academia and industry. This knowledge is avail-
able to learners in the form of a large collection of online re-solvable examples and submodels of systems
components and physical phenomena.

References
Multidisciplinary systems
R. H. Cannon: Dynamics of Physical Systems. McGraw-Hill, New York 1967

H. E. Koenig, Y. Tokad, H. K. Kesavan: Analysis of Discrete Physical Systems. McGraw-Hill, New


York 1967

S. H. Crandall, D. C. Karnopp, E. F. Kurtz, D. C. Pridmore-Brown: Dynamics of Mechanical and


Electromechanical Systems. McGraw-Hill, New York 1968

J. L. Shearer, A. T. Murphy, H. H. Richardson: Introduction to System Dynamics. Addison Wesley,


Reading 1971

F. E. Cellier: Continuous System Modeling. Springer-Verlag, New York 1991

D.C. Karnopp, D.L. Margolis, R.C. Rosenberg: System Dynamics: A Unified Approach. John Wiley,
New York 1990 (2nd ed.)

M.M. Tiller: Introduction to Physical Modeling with Modelica. Kluwer Academic Publishers 2001

P. Gawthrop, L. Smith: Metamodelling: Bong graphs and dynamic systems. Prentice Hall 1996.

R. Bjrke: Manufacturing Theory. Tapir 1995.

R. Isermann: Mechatronische Systeme. Grundlagen. Springer 1999.

F.T. Forbes: Modeling and Analysis of Engineering Systems, a Unified Graph-Centered Approach,
Marcel Dekker 2001.

R.A. Layton: Principles of Analytical System Dynamics. Springer 1998.


CONTENTS 5

D.K. Miu: Mechatronics. Electromechanics and Contromechanics. Springer 1993.

D. Rowell, D.N. Wormley: System Dynamics: An Introduction. Prentice Hall, 1997.

J.L. Shearer, B.T. Kulakowski, J.F. Gardner: Dynamic Modeling and Control of Engineering Systems.
Prentice Hall, 1997.

R.L. Woods, K.L. Lawrence: Modeling and Simulation of Dynamic Systems. Prentice Hall, 1997.

I. Cochin, W. Caldwallender: Analysis and Design of Dynamic Systems. Addison-Wesley, 1997 (3rd.
ed.)

W.J. Palm III: Modeling, Analysis, and Control of Dynamic Systems. John Wiley & Sons, 1999.

Electrical, magnetic and electronic systems


D.H. Sheingold: Nonlinear Circuits Handbook. Designing with Analog Function Modules and ICs.
Analog Devices, Norwood MA 1974

Fano, R.M., Chu, L.J.: Fields, Energy, and Forces. Wiley, New York 1959.

M.H. Rashid: SPICE for Circuits and Electronics Using PSpice. Prentice-Hall Int. Editions, Engl. Ciffs
1990.

Slemon G.R.: Magnetoelectric Devices. Transducers, transformers, and machines. John Wiley, New
York 1966.

R. Kielkowski: Inside SPICE. Overcoming the Obstacles of Circuit Simulation. McGraw-Hill, New York
1994.

J.A. Connelly, P. Choi: Macromodeling with SPICE. Prentice Hall, Engl. Cliffs 1992.

K.H. Muller: A Spice Cookbook. Intusoft 1991.

M.H. Rashid: Power Electronics. Circuits, Devices, and Applications. Prentice-Hall Int. Editions, Engl.
Ciffs 1988.

C.A. Desoer, E.S. Kuh: Basic Circuit Theory. McGraw-Hill 1969

J. Vlach, K. Singhal: Computer Methods for Circuit Analysis and Design. Van Nostrand Reinhold 1994
(2nd ed.).

N. Mohan, T.M. Undeland, W.P. Robbins: Power electronics. Converters, Applications, and Design.
Wiley 1995 (2nd ed.).

G. Schweitzer, H. Bleuler, A. Traxler: Active Magnetic Bearings. VDF 1994.

Mechanical systems
P.E. Nikravesh: Computer-Aided Analysis of Mechanical Systems. Prentice-Hall 1988

E.J. Haug: Computer Aided Kinematics and Dynamics of Mechanical Systems. Allyn and Bacon 1989.

R.D. Klafter, T.A. Chmielewski, M. Negin: Robotic Engineering. Prentice-Hall 1989.

J.L. Meriam, L.G. Kraige: Engineering Mechanics. Statics. Dynamics. Wiley 1998 (4th ed.).

R.L. Norton: Design of Machinery. McGraw-Hill 1999.

F.S. Tse, I.E. Morse, R.T. Hinkle: Mechanical Vibrations. Theory and Applications. Prentice-Hall
1978.

SimMechanics, The MathWorks, Natick MA, 2002.

C.A. Rubin: Working Model. Addison-Wesley 1995.


6 CONTENTS

J.J. Craig: Introduction to Robotics. Mechanics and Control. Addison-Wesley 1989 (2nd ed.).
DUBBEL Handbook of Mechanical Engineering. Springer 1994.
H. Mayr: Virtual Automation Environments. Design, Modeling, Visualization, Simulation. Marcel
Dekker 2002.

Electro-mechanical systems
A. Lenk, G. Pfeifer, R. Wertschtzky: Elektromechanische Systeme. Mechanische und akustische Net-
zwerke, deren Wechselwirkungen und Anwendungen. Springer 2001.
W. Leonhard: Control of Electrical Drives. Springer 1985.
P. Krause, D. Wasynczuk, S. D. Sudhoff: Analysis of Electric Machinery. IEEE Press 1995.
H. Gross: Electrical Feed Drives for Machine Tools. Wiley 1983.
P.C. Krause: Analysis of Electric Machinery. McGraw-Hill 1987.
H.H. Woodson, J.R. Melcher: Electromechanical Dynamics. Part I: Discrete Systems. Part II: Contin-
uum Electromechanics. Part III: Elastic and Fluid Media. Wiley 1968 [repub. by Krieger 1990].
J. Meisel: Principles of Electromechanical-Energy Conversion. McGraw-Hill 1966.
D.C. White, H.H. Woodson: Electromechanical Energy Conversion. Wiley 1959.
T. Kenjo: Stepping motors and their microprocessor controls. Clarendon Press 1992.
R.T. Smith: Analysis of Electrical Machines. Pergamon Press 1982.
S.A. Nasar, L.E. Unnewehr: Electromechanics and Electric Machines. Wiley 1979.
J.R. Melcher: Continuum Electromechanics. MIT Press 1981

Fluid and thermal systems


T. J. Viersma: Analysis, Synthesis and Design of Hydraulic Servosystems and Pipelines. Elsevier,
Amsterdam 1980

J.H. Lienhard: A Heat Transfer Handbook. Prentice-Hall, Inc., Englewood Cliffs, N.J., 1981.
Y.A. Çengel: Thermodynamics. An Engineering Approach. McGraw-Hill, New York 1994 (2nd ed.).
H.E. Meritt: Hydraulic Control Systems. Wiley 1967.
J. Watton: Fluid Power Systems. Modelling, simulation, analog and microcomputer control. Prentice
Hall 1989.
Module 1

Introduction to dynamics

Module units
1.1 Modeling and simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-1
1.2 Role of simulation in engineering . . . . . . . . . . . . . . . . . . . . . . . . . 1-2
1.2.1 Design problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-2
1.2.2 Design process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-4
1.2.3 Design descriptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-4
1.2.4 Design levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-6
1.2.5 Design methodologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-7

Module overview. Welcome to the course part focused on system dynamics. It is well recognized
that computer-assisted modeling and simulation are indispensable for understanding dynamics of existing
systems as well as for designing new systems. This module attempts to explain why is it so.

1.1 Modeling and simulation


The behavior of real physical systems which engineers must design, analyze, and understand is controlled
by the flow, storage, and interchange of various forms of energy. In almost all cases, real systems are
extremely complex and may involve several interacting energy phenomena, such as electromagnetic fields,
heat and fluid flow, or nuclear reaction.
Simulation or analysis of a dynamic system always involves forming a model made up of basic building-
blocks that are idealizations of the essential physical phenomena occurring in real systems. An adequate
model of a particular physical device or system will behave approximately like the real system. The
best system model is the simplest one which yields the information necessary for engineering action or
decision.
It is important that the learners appreciate the advantages of simplified ideal models without over-
simplifying their picture of the real physical world. An approximate answer to an engineering problem
is often the best answer if the degree of approximation is known and a more precise answer is not really
needed. The cost of precision is often exceedingly high, since the model complexity of the conceptual
models needed to represent systems usually increases very rapidly with increased requirements for accu-
racy. Hence efforts to attain high precision should be made only when the cost in time, money, and effort
can be justified.
Of the many fundamental building blocks which the engineer must use in forming models of particular
dynamic systems, we shall concentrate in this part of the course on simple physical elements which store,
dissipate, and transform energy. By proper use of these idealized elements, we will be able to model
any physical system, linear or nonlinear. The extent to which the behavior of the model represents the
actual behavior of the real physical system depends on the sophistication and completeness of the model
and on the degree of insight and judgment with which the model is formulated. When the engineers are
confronted with a dynamic system problem, they must first recognize that the problem is one to which
they can apply the principles developed in this course.

1-1
1-2 MODULE 1. INTRODUCTION TO DYNAMICS

We shall see that there are many striking similarities between mechanical, electrical, magnetic, fluid,
acoustic, and thermal systems, which will permit a unified study of the system behavior. A uniform
terminology and symbolism applicable to all of the physical systems considered will be developed. The
integration of subject matter in this course is designed to allow the learners to recognize that a large
class of problems can be considered in a similar and systematic fashion.
In the next step, the engineers must define the system to be considered (should the ambient-temperature
variation be considered an input? is the inertia of a connecting shaft important in this situation? etc.).
They then may describe the system by means of the various physical elements or more complex submodels
which we will consider. After this, they must investigate the energetic interactions between these system
parts.
The next job is to establish a mathematical model of the system to be analyzed. This involves the
identification and idealization of individual system components as well as identification and idealization
of their interconnection. The mathematical statement of the governing relationships between system
variables is called the formulation problem. Interconnection of the elements imposes constraints on the
variation of system variables, and the convenient way of specifying these constraints is by a mathematical
statement of the way in which the various variables are related.
To investigate the dynamic behavior of the system, this set of equations must be solved. The initial
state of the system must also be specified by a set of initial conditions. The system equations must be
solved for the desired response under the specified input signal and initial conditions. There are numerous
approaches to the solution of the system equation, and the choice of a method or software will depend
on the problem at hand.
After obtaining a solution for the response, the engineers have another major job. The tasks of initial
modeling and final analyzing or designing are functions which clearly distinguish the engineer from the
mathematician. The engineers must check their solution (is it correct dimensionally? does it correspond
to physical reality? does it check for simplified situations which can be easily analyzed? etc.). Their
interest in the whole matter of dynamic systems is to eventually analyze to determine whether a certain
performance is obtained and is satisfactory, or more generally, he must design the system so that it will
meet certain performance specifications.
In summary, our approach to a dynamic system problem can be enumerated as follows:

1. define the system and its components,

2. formulate the mathematical model,

3. determine the system equations,

4. solve for the desired output,

5. check the solution,

6. analyze or design.

1.2 Role of simulation in engineering


1.2.1 Design problem
Fig. 1.1 illustrates the life-cycle of an engineering system. The life cycle begins from design of the system,
the starting point of which is system specification based on marketing as well as technical, economic,
environmental and other requirements. The specification defines the correct system in terms of all critical
system properties within the assumed range of operating conditions throughout the system complete life
cycle up to the system disposal.
The most rudimentary design activity, the nominal design, aims at achieving the requirements on
the correct system under the nominal manufacturing and operating conditions. In reality, however,
these conditions are never exactly nominal due to manufacturing inaccuracies, variations of the system
environment, aging of the system, etc. Thus the design process consists of a variety of additional design
activities with different but complementary objectives.
An important role is played in the design process by tolerance design which takes into consideration the
fact that parameters of components the systems are made of as well as parameters of the environment
in which the systems operate deviate from their nominal values but still stay within the admissible
1.2. ROLE OF SIMULATION IN ENGINEERING 1-3
 
SYSTEM SPECIFICATION
 
?
DESIGN OF SYSTEM - DESIGN OF TESTS
?
DESIGN VERIFICATION 

?
MANUFACTURING

?
OUTPUT TESTING 

?
SYSTEM OPERATION

?
MAINTENANCE 

?
SYSTEM OPERATION

Figure 1.1: Life-cycle of an engineering system.

tolerances. There are two basic approaches to the problem: worst-case and statistical tolerance design.
Worst-case design aims at preventing system failure even in the case of the most adverse combinations
of the parameter deviations. Very few systems, however, are designed for 100% production yield, i.e.,
that all 100% ‘copies’ of the system are correct at the output of the production process. The reasons
are economic one: it is cheaper to dump or repair a part of the production than to utilize components
with tolerances more narrow. The aim of statistical tolerance design is to achieve the required system
production yield with components of given statistical distribution of their tolerances.
Of course, a system may fail also if one or more of the system components fail, i.e. if the properties
of these components get outside their admissible tolerances. The probability that the system would
not fail in this way during certain time period and under certain operating conditions is called system
reliability. The objective of reliability design is to meet the requirements on the system reliability taking
into consideration the limited reliability of the system components.
Design for testing aims at designing systems in such a way that they could be tested easily, some
systems are even designed to be self-testing. The design for testing is complemented by design of tests.
As Fig. 1.1 suggests, different tests are needed for design verification, for testing at the output of the pro-
duction, or for testing of systems during their in-service operation and maintenance. Such experimental
testing is increasingly carried out using computer-controlled instruments.
The objectives of system testing can be of different levels of depth:

• fault detection recognizes only the system failure occurrence

• fault localization reveals in which of the system components the failure occurred

• fault diagnosis indicates the reason for the failure

There are also other design activities, like design for producibility, for maintainability, for reproducibil-
ity, for built-in obsolescence, for self-destruction, etc.
In all these activities computer-assisted modeling, simulation and analysis makes an indispensable
tool today. Used to explore and predict dynamic behavior of dynamic systems, it plays a key role in
enhancing their quality and reliability. Simulation also decreases the need for costly and time consuming
1-4 MODULE 1. INTRODUCTION TO DYNAMICS

DESIGN DESCRIPTIONS:
behavioral topological geometrical
DESIGN LEVELS:

conceptual

functional
DESIGN SPACE
physical

technological

Figure 1.2: Design space

production of real prototypes and for their experimental testing. Many of these activities cannot be
performed experimentally at all, like the tolerance or reliability design, for example. Last but not least,
by speeding up the design process, simulation allows for acquiring higher profits by introducing new
products into the market faster.

1.2.2 Design process


The process of designing is usually an iterative analysis. The results of one analysis point toward a
change in the system which may improve the performance; another analysis is done on the new system
to determine performance, etc. The design aspects distinguish the engineer from the scientist. The
engineer is ultimately interested in building a system which will perform a useful function for the benefit
of mankind.
To manage the designing of an engineering system, the process is usually carried through several
design levels starting from the conceptual up to the technological design level. At the same time, several
different design descriptions of the system are used in different phases of the process. In addition to
it, if the system is complex, it is imagined as decomposed into numerous design modules which can be
designed to a certain degree autonomously. Also each module can be considered as decomposed into
several submodules, and such a decomposition can proceed through several levels of module complexity.
Thus the process of designing an engineering system can be imagined as carried through a three-
dimensional design space the typical settings of which are given in Fig. 1.2. The fundamental design
levels are shown along the vertical border of the design space whereas the basic design descriptions are
indicated along its horizontal border. The module complexity scale is not shown in the figure – it is
assumed to be associated with the axis perpendicular to the plane of the figure.
The decomposition of a system into design modules or subsystems is a matter of engineering judgment.
The modules may represent both separable parts from which the system is actually assembled (like various
mechanical members, actuators, electronic circuits, hydraulic elements, etc.) as well as just some effects
which are not separable from the parts (like friction or flexibility, for example).
To illustrate the process, let us take the example of a copying lathe given in Fig. 1.3a. In Fig. 1.3b,
the lathe is shown disintegrated into several modules: the lathe bed with the slide, the hydraulic cylinder,
the spool valve, the stylus following a master, the tool manufacturing the workpiece and, as a separate
part, the surfaces of the cylinder container and of the slide between which a friction effect takes place.
Fig. 1.3a gives an example of a geometric description of the copying lathe in the form of its cutaway view.

1.2.3 Design descriptions


Each of the design descriptions represents a system or module model of certain degree of abstraction and
idealization. The geometric description of a system or module expresses its geometrical shape, dimensions
and position. In the past, the geometric description used to be recorded in the form of a blue print whereas
today it is usually stored in an electronic form which may represent not only the traditional engineering
drawings of the module, but also its wireframe, solid, surface or some other geometric model.
1.2. ROLE OF SIMULATION IN ENGINEERING 1-5

(a)

C D

x
(b)
. D

A S
C

B
W N
f

(c)

Figure 1.3: (a) Copying lathe, (b) lathe decomposition, (c) multipole model.
1-6 MODULE 1. INTRODUCTION TO DYNAMICS

The behavioral description characterizes external dynamic interactions of a module in terms of the rela-
tionship between the interactions ignoring completely the module internal structure, e.g., it characterizes
the module just as a ‘black-box’. The model, which the behavioral description of a module represents,
may be static or dynamic, linear or nonlinear, time-variable or time invariant, lumped-parameter or
distributed-parameter, etc. The description may be in the form of various characteristics obtained by
experimenting with the real module or in the form equations, transfer functions, state-transition tables,
etc., based on a hypothesis about the physical effects taking place within the module. The experimental
and hypothetical approaches are usually combined.
On the other hand, to represent mutual interactions between modules in a system, or between submod-
ules in a module, is the aim of the topological description. Any module geometrical shapes, dimensions
and positions are meaningless in this description, only the coincidences of the module interactions are
relevant. The topological descriptions are usually portrayed graphically in the form of diagrams in which
modules or their functions are represented by some symbols and the module interactions by line segments
interconnecting the symbols.
Fig. 1.3c gives an example of a topological description of the copying lathe. Besides the standardized
symbols for a spool valve and a hydraulic cylinder, we can see there a damper bf modeling the friction
between the cylinder and the slide, the combination of damper bW and spring kN modeling the dynamics
of the manufacturing process, the actuator of velocity ṙ enforced to the stylus by the rotating master,
the square symbol representing inertial mass m of the movable parts, and the symbol S representing
there for the oil pressure supply. The interactions between the modules are denoted in Fig. 1.3c by
small circles. Thus the letters A and B denote interactions carried out by mechanical interlinking of the
modules whereas C and D denote interactions between the module fluid inlets.

1.2.4 Design levels


The uppermost design level is committed to the conceptual design of the system. It results in an overall
architecture and fundamental functions of the designed system. Also decisions of which the functions
will be implemented in hardware, either analog or digital, and which of them will be implemented in
software. The behavioral description at this level corresponds directly to the introductory specification of
the system. The topological description, usually in the form of a block diagram, indicates at this level the
fundamental ‘causes-and-effects’ within the system. An example of such a block diagram for the copying
lathe is shown in Fig. 1.4a. The geometrical description displays the system ‘floor-plan’, i.e. its overall
shape and dimensions.
At the functional design level, the system architecture and functions are refined into more detail. The
modules at this level are assumed to interact in the form of physically dimensionless continuous-time
continuous-level signals, or signals discretized in time or in level or both. The module interactions are
governed by algebraic rules. The topological description used at this design level is mostly in the form
of block diagrams with blocks characterized by equations, transfer functions, tables, etc. Typically, the
design of control strategy and control architecture, either analog or digital, is a matter of functional-level
design.
An example of a block diagram showing control of the copying lathe is given in Fig. 1.4b. Two inputs
affecting the control are considered in the diagram: stylus displacement r, and external forces caused
by load and friction F . Output y is the machine-tool displacement, ε stands for control error and s is
the Laplace operator. Velocity gain K and load stiffness c are parameters of the lathe. Note that the
structure of this block diagram has nothing in common with the geometric structure of the real lathe
shown in Fig. 1.3a. To set up this diagram, equations based on the physical laws governing the lathe
operation must be formulated first, then simplified for the control synthesis purposes, and only after these
two steps converted into the diagram.
The physical design level is concerned about hardware implementation of the system architecture
and functions. Also ‘parasitical’ physical effects (like unwanted vibrations or heat generated during
system operation) are considered there. Design modules take the form of real system components or
physical effects and their mutual energetic interactions are expressed in terms of physical quantities. The
topological description at the physical level takes on the form of a multipole diagram. An example of
such a multipole diagram for the copying lathe is shown in Fig. 1.3c.
Note the fundamental differences between block and multipole diagrams:

• Each of the line segments interconnecting blocks in a block diagram represents an unidirectional
transfer of one mathematical variable. The line segments interconnecting multipoles in a multipole
1.2. ROLE OF SIMULATION IN ENGINEERING 1-7

Figure 1.4: Copying lathe block diagrams at (a) the conceptual and (b) functional design level.

diagram represent bilateral transfer of energy between system modules via electrical wires, fluid
lines, shafts or other mechanical links, etc. Each of the line segments is associated with two physical
quantities the product of which is a physical dimension corresponding to power.
• The mutual interactions of blocks follow algebraic rules whereas the multipole interactions are
governed by physical laws of energetic balance and geometric compatibility.
• The block diagram is just a graphical representation of equations. To set up this diagram these
equations must be formulated first and then converted into the diagram. The structure of this
diagram has nothing in common with the structure of the real system which the equations charac-
terize. The topological structure of the multipole diagram corresponds directly to the geometrical
structure of the real system. Thus setting up of this diagram can be based simply on inspection of
the real system.
At the technological design level the assembly or another manufacturing process for the system and
for its components is designed.

1.2.5 Design methodologies


Obviously, the aim of the design process is to proceed through the design space from its upper-left
corner, representing the system specification, up to its lower-right corner, corresponding to the detailed
technological documentation. The actual route through the design space corresponds to the chosen design
methodology.
There are two extremes in design methodologies: bottom-up and top-down. The bottom-up method-
ology, consisting of consecutively building the system from available components based on a sequence of
design decisions, does not guarantee achieving an optimal system design. In fact, it does not guarantee
that the system specification will be met at all. The top-down methodology is the most preferable, as it
leads to optimal results. The down-going design routes, however, are subjected to constraints imposed
by the lower design levels. Thus the actual design routes consist inevitably of many horizontal as well as
vertical loops corresponding to the individual design iterations.
1-8 MODULE 1. INTRODUCTION TO DYNAMICS
Module 2

Rectilinear mechanical systems

2.1 Rectilinear systems and their surroundings


To define the rectilinear system of our interest, we usually begin the investigation of the system dynamic
behavior with drawing a sketch of the system geometric configuration showing dimensions and
positions of the system bodies. To determine directions of the body motions and of the related forces,
we associate the system sketch with a coordinate axes.
Fig. 2.1 shows, for example, a sketch of two cars in the rectilinear motion along a straight horizontal
road. A car is towing another car using a towing rod connected between the car hooks A and B. The
assumed positive direction of the car motions is in agreement with the direction of the x axis.

Figure 2.1: Car towing.

Bodies of the cars constitute together with the body of the rod a mechanical rectilinear system.
Behavior of the system is influenced by the system surroundings formed by the wind blowing opposite to
the car motions as well as by the friction between the car tyres and the road surface.

Figure 2.2: Wheel suspension system.

Fig. 2.2 was sketched for the purpose of the dynamic analysis focused on the behavior of a car-wheel
suspension. When the car travels along a road, the wheel tyre copies potholes in the road surface. The
wheel is thus driven up or down in the vertical direction along the axis y. In this case, the rectilinear
system under consideration consists of four bodies: the spring, the shock absorber, the wheel and the
quarter of the car body. The system excitation by the road surface and the gravitational attraction of
the wheel and the quarter-car body forms the system surroundings in this case.

2-1
2-2 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

2.2 Rectilinear interactions


2.2.1 Rectilinear translation
Rectilinear motion of a point is defined as the motion of that point along a straight line. Translational
motion of a body can be either rectilinear or curvilinear as shown in Fig. 2.3. All points of a body in
rectilinear translation move in parallel straight lines. This is in contrast to curvilinear translation
of bodies the points of which move on congruent curves.

Figure 2.3: Body in (a) rectilinear and (b) curvilinear translation.

If a body can be considered rigid its rectilinear motion is completely specified by the motion of any
one point in the body, since all points have the same motion. A rigid body is a body whose changes in
shape are negligible compared with the overall dimensions of the body or with the changes in position of
the body as a whole. If each of the cars shown in Fig. 2.1 is considered as a rigid body, motions of the
cars can be represented, for example, by the points A and B denoting the cars’ rod hooks.
Absolute rectilinear motion of a given point is defined as the motion of that point relative to
a fixed point in space considered as a reference frame for the motion. Because of the difficulty of
providing a real fixed point in space from which to make measurements, the concept of absolute motion
is strictly hypothetical, but nevertheless very useful.
A reference frame fixed to the earth is often a reasonable approximation. However, when dealing with
the motion of long-range missiles, for example, the earth cannot be considered to be a nonaccelerating
reference and velocities are then measured with respect to the ‘fixed’ stars instead.
Fig. 2.4 shows a geometric configuration of points A and B representing the car hooks in Fig. 2.1.
Motion of the points is constrained to the rectilinear translation along the x-axis ‘painted’ on the reference
frame. The position xA (t) of point A is defined as the distance of this point from the origin 0 chosen
on the axis x. In the figure, xA (t) is shown negative whereas the position xB (t) of point B is positive.
Positions are measured in meters [m].

Figure 2.4: (a) Positions and (b) velocities of points in rectilinear motion.

The absolute velocity ẋA (t) [m/s] and the absolute acceleration ẍA (t) [m/s2 ] of point A is by
definition
d dẋA d2 xA
ẋA (t) = xA and ẍA (t) = =
dt dt dt2
We shall depict the absolute velocities in drawings of geometric configurations by empty-head arrows
as shown in Fig. 2.4. Displacements, velocities and accelerations are taken as positive if they correspond
to a motion in the chosen positive direction of the related coordinate axis.
We shall often be concerned with the relative motion of two points. The relative displacement xBA ,
relative velocity ẋBA and relative acceleration ẍBA of point B with respect to point A is at any instant
xBA = xB − xA , ẋBA = ẋB − ẋA , and ẍBA = ẍB − ẍA (2.1)
Position, velocity or acceleration of a point with respect to another point can be measured by a device
applied to the two points “across” the system without braking into it. Velocity can be thus considered as
2.2. RECTILINEAR INTERACTIONS 2-3

an across variable, position as an integrated across variable, and acceleration as a differentiated across
variable.

2.2.2 Forces
Force is the action on a body characterized by its magnitude, by the direction of its action, and by its
point of application. Forces are measured in physical units called newton [N]. A force tends to move a
body in the direction of its action.
Forces acting on a body are classified as either contact forces or body forces. Contact forces are
generated through direct physical contact between two bodies or between a body and a fluid. Body
forces are those applied by remote action, such as gravitational, electrostatic or electromagnetic forces.
Actually both contact and body forces are distributed forces as the contact forces are in fact distributed
over the cross-section area of the contact. In the system shown in Fig. 2.1, the forces of the wind acting
on the cars are distributed over the car surfaces, while the forces in the towing rod are distributed over
the rod cross-section area. If a rigid body is constrained to rectilinear translation all the forces acting on
the body can be considered as concentrated forces acting at a point.
To take a complete account of all forces acting on a body, it is essential to construct the free-body
diagram by isolating the body completely from the rest of the system and its surroundings. This diagram
consists of a closed outline of the external boundary of the body. The diagram shows all the external
contact and body forces applied to the body.

Figure 2.5: (a) Free-body diagram, (b) concentrated applied force.

Fig. 2.5a shows the free-body diagram of a towed car, a part of the rectilinear system introduced in
Fig. 2.1. Ft is the force exerted on the car by the towing rod, Fw represents the distributed force of wind
acting on the car surface, and Ff is the friction force acting on the car wheels due to the roughness of
the road.
As it is indicated by full-head arrows, we have chosen the following convention for the orientation of
forces applied to a body. The assumed positive direction of all applied forces is in agreement with the
direction of the reference axis. Values of those applied forces that tend to accelerate the body in this
direction are positive, if an applied force tends to decelerate the body its value is negative.
When we are dealing with an external action of a force on a rigid body in rectilinear motion, the force
may be applied at any point of the body without changing its effect on the body as a whole. Thus, if the
car shown in Fig. 2.5a is considered as a rigid body, all three forces applied on it can be replaced by one
concentrated force Fa acting for example at point A as shown in Fig. 2.5b. Obviously,

Fa = Ft + Fw + Ff (2.2)

A force F can be viewed as the time rate of the linear momentum µ [N.s] (called also impulse)
acting in the same direction, i.e.
d
F (t) = µ = µ̇ (2.3)
dt
or
Z t
µ= F (t) dt + µ0 (2.4)
0

where µ0 is the momentum at t = 0.


Forces can be measured directly by means of a calibrated spring, for example. To measure the force
at a point directly, we must severe the system at that point and insert the measuring device between the
two resultant system sections. Therefore, forces are thought of as being applied “through” the measuring
device and hence can be considered as belonging to the family of through variables.
2-4 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

2.2.3 Power, energy and work


The mechanical power P delivered to a moving point by the force F which acts at that point in the
direction of the positive velocity v is the product of the velocity and the component of force which lies
in the direction of the velocity. If the force F and the velocity v are collinear,

P(t) = v(t) F (t) (2.5)

Power P is defined as the rate of flow of energy E or work W so that


d d
P(t) = E= W (2.6)
dt dt
In agreement with our convention for the orientation of applied forces, we will assume that energy is
consumed in the body if the signs of the body absolute velocity and applied force are the same (both
signs are positive or both are negative).
The mechanical work W done by one body on another through a point of connection is the time
integral of the power: Z tb
Wab = v F dt (2.7)
ta

where ta and tb are the beginning and end of the time interval over which we desire to compute the work
done.
The mechanical energy E supplied to a body is the sum of the work done on the body by all forces
acting on it
Xn
E= (Wab )k (2.8)
k=1

2.3 Composition of rectilinear interactions


2.3.1 Composition of rectilinear motions
Let us consider again Fig. 2.4 showing a geometric configuration of points A and B constrained to the
rectilinear translation along the x-axis. According to (2.1), the relative displacement xBA , relative velocity
ẋBA and relative acceleration ẍBA of point B with respect to point A is at any instant

xBA = xB − xA , ẋBA = ẋB − ẋA , and ẍBA = ẍB − ẍA (2.9)

These relations represent geometric constraints which are placed on the collinear rectilinear mo-
tions of points due to geometric continuity of space. This is just a demonstration of the postulate of
compatibility in rectilinear translation.

Re-solvable example
Composition of motions

2.3.2 Composition of forces


Fig. 2.6a gives an example in which point A constrained to rectilinear motion along the x-axis is acted
on by forces F1 (t) and F2 (t) collinear with x. The assumed positive direction of the forces is indicated
by the full-head arrows. According to Newton’s third law of action and reaction, the forces acting on
a point are in equilibrium, i.e., their resulting effect is zero. Thus,

F1 (t) + F2 (t) = 0, or F1 (t) = −F2 (t) (2.10)

This shows that the forces F1 (t) and F2 (t) are of the same magnitude but of the opposite direction at
any instant. If F1 is considered as an action force, F2 then forms the reaction force with respect to F1 ,
and vice versa.
The action of a force on a body can be separated into two effects, external and internal. Forces
external to a body are of two kinds, applied forces and reactive forces. The effects of applied forces
internal to the body are the resulting internal stresses and strains distributed throughout the material of
2.3. COMPOSITION OF RECTILINEAR INTERACTIONS 2-5

Figure 2.6: Balance of forces acting on (a) a point, (b) on a body, (c) on an ideal link, (d) on a configuration
of ideal links.

the body. The relation between internal forces and internal strains involves the material properties
of the body and is studied in strength of materials, elasticity, and plasticity. In mechanics of rigid bodies,
concern is given only to the net external effects of forces. Fig. 2.6b shows the external force Fa applied
on a car and the corresponding reaction force of the car Fc = Fa .
The principle of transmissibility states that an external force applied at a point of a rigid body
constrained to rectilinear motion may be considered to be transmitted or to flow unchanged through the
body to any other point of the body.
Fig. 2.6c shows points A and B moving along the x-axis. The points, which are interconnected by an
ideal link, are acted on in the x-direction by forces F1 and F2 external to the link. The ideal rectilinear
link is by definition absolutely rigid and has no mass. The link is characterized by the relations
F1 (t) + F2 (t) = 0 (2.11)
ẋA (t) = ẋB (t) (2.12)
The relation (2.12) is identical with (2.10). This indicates that due to the principle of transmissibility
the force equilibrium holds also if the forces F1 (t) and F2 (t) applied to the endpoints of an ideal rectilinear
link. The second relation (2.12) follows from the assumption that there are no link deformations due to
forces acting on the link. As the link is massless it is not subjected to any inertial forces and it neither
accumulates, nor dissipates any energy.

Figure 2.7: (a) Applied and reaction forces. (b) Reaction forces flowing through the system. (c) Reaction
forces in the pair of scales.

This example may be generalized for any number of points interconnected by ideal links. Fig. 2.6c
shows three points A, B and C acted on by forces F1 , F2 and F3 , respectively. The points are linked to
2-6 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

point D. All the points are in rectilinear translation along the x-axis. The three forces are in equilibrium
and the point velocities are identical, i.e.,

F1 (t) + F2 (t) + F3 (t) = 0 (2.13)


ẋA (t) = ẋB (t) = ẋC (t) = ẋD (t) (2.14)

Fig. 2.7a shows two cars interconnected by a towing rod. The cars are considered as rigid bodies and
the rod as an ideal rectilinear link. The assume positive orientation of the car reaction forces Fc1 and Fc2
shown within the car outlines opposes the x-direction. The car reaction forces, which can be imagined
as flowing through the two-car assembly, are in balance, i.e.

Fc1 + Fc2 = 0 (2.15)

To prove (2.15), let us consider Fig. 2.7b showing free-body diagrams of the cars and towing rod. The
forces denoted outside the cars and the rod are applied forces the assumed positive direction of which is
the x-direction. Let us assume that

Fa1 = Fr1 + Fo1 and Fa2 = Fr2 + Fo2

where Fr1 and Fr2 are forces applied on the cars by the towing rod, and Fo1 and Fo2 are the other external
forces applied on the cars. When submitting these relations into (2.15) after taking into consideration
that Fc1 = Fa1 and Fc2 = Fa2 we obtain

Fr1 + Fo1 + Fr2 + Fo2 = 0

The forces Fr1 and Fr2 acting on the rod are in balance, and the other external forces Fo1 and Fo2 applied
to the system of rigid bodies must be in balance too.
The reaction forces in the vertical links of the pair of scales shown in Fig. 2.7c must be balanced, i.e.

Fa + Fb + Fc = 0

This conclusion can be viewed as a demonstration of the postulate of continuity in rectilinear


translation.

Re-solvable example
Static equilibrium of forces

2.4 Rectilinear inertia of mass


2.4.1 Pure inertor
Components of mechanical systems having mass resist changes of their velocity due to inertia forces
acting on the mass. Let us consider a body of a constant mass m [kg] in a rectilinear motion of absolute
acceleration a [m/s2 ]. According to Newton’s second law, the inertia force F [N] acting on the mass
in the opposite direction to the mass acceleration is
d
F (t) = m a(t) = m v(t) (2.16)
dt
where v [m/s] is the absolute velocity of the mass with respect to a nonaccelerating reference frame.
Using the definition of force (2.3), Newton’s second law (2.16) can be expressed also as

µ(t) = m v(t) (2.17)

where µ is the linear momentum acting on the mass. By definition, µ = 0 when v = 0. Thus
Z t
mv = F (t) dt + µ(t0 ) (2.18)
t0

The inertia effects encountered in rectilinear translation of bodies can be modelled by the idealized
physical element called pure inertor. The inertor ignores, however, damping, spring or any other
2.4. RECTILINEAR INERTIA OF MASS 2-7

Figure 2.8: (a) Symbol of inertor. (b) Momentum vs. velocity characteristic of ideal inertor. (c) Sliding
body, (d) its model.

energy effects in bodies. Our symbol for the inertor is shown in Fig. 2.8a. The symbol consists of a
square representing the mass and of a short line segment called a pin. The pin denotes the pole of the
inertor which represents the inertor energy interactions with the rest of the system.
The pure inertor is associated with two variables: the inertor force F representing the inertia force
of the modelled mass, and the inertor velocity v representing the absolute velocity of the mass. The
assumed positive orientation of F and v with respect to the inertor symbol is indicated in Fig. 2.8a by
the full-head and empty-head arrow, respectively.
Dynamic behavior of the pure inertor is completely described by the constitutive relation

µ = f (v) (2.19)

where f (·) is a function. A pure inertor characterized by the linear constitutive relation (2.17) or (2.16),
where the mass m is constant, is called the ideal inertor. The relationship (2.17) is plotted in Fig. 2.8b.
From the theory of relativity we know that mass of a given body increases with the body velocity
approaching the velocity of light. As the present considerations are limited to velocity levels small
compared with the velocity of light, Newton’s second law can be applied. No other nonlinear inertia
effects are known. In some engineering problems, however, the mass of a body changes with time.
Fig. 2.8c shows the free-body diagram of a rigid body sliding with a negligible friction along a hori-
zontal surface in the x-direction. Let us assume that point A was chosen to represent the body rectilinear
motion as well as the site of application of all the external forces Fext acting on the body. Such a point
can be considered as the body mass point, i.e. a point of the body associated with the lumped body
mass constrained to the rectilinear motion.
Fig. 2.8d gives the dynamic model of the sliding body in the form of a multipole diagram. Point A of
the real body moving in the x-direction as shown in Fig. 2.8c is represented in the diagram by the node
Ax . Diagram nodes represent sites of energy interaction between bodies.
This is in agreement with our convention about the orientation arrows for reactions to external forces
applied to bodies. The empty-head arrow, which denotes the assumed positive orientation of the inertor
velocity v, is directed towards the inertor pole away from the reference frame.
As we assume that the body is not subjected to any other effects except its own inertia, there is only
one physical element in the diagram – the inertor – which is attached to node Ax . If no external force
Fext is acting on the body in this example, the inertia force Fi is zero and the body velocity ẋA is also
zero or constant.
The inertor is in fact a two-pole physical element. But its second pole, with respect to which the
inertor velocity should be ‘measured’, must be always attached to the reference frame. For this reason,
the second pole need not be shown in the inertor symbol just to simplify the diagrams.
2-8 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

2.4.2 Inertia energy


From (2.16) it is obvious that the power entering an ideal inertor at a given instant is

P(t) = v(t) F (t) = m v(t) v̇(t) (2.20)

If P > 0, energy is being stored in the inertor, and if P < 0, the stored energy is being retrieved from
it. Which of these two states is occurring depends both on the relative velocity and acceleration at the
moment.
Integration of (2.20) gives the expression for the energy accumulated in the ideal inertor
1
E= m v2
2
which is called rectilinear kinetic energy. The energy is represented by the area in Fig. 2.8b between
the µ axis and the µ-v characteristic.
We can see that this energy depends directly on the velocity acquired by the inertor, but not on its
force. Hence the inertor stores energy by virtue of its velocity. The inertor energy is always positive and
does not depend on the sign of either its velocity or acceleration.

2.5 Prescribed rectilinear interactions


2.5.1 Prescribed velocities
If a body is driven by a prescribed velocity v(t) [m/s], the actuator driving the body can be modelled
by the pure source of velocity. This idealized physical element is characterized by the constitutive
relation
v(t) = f (t) (2.21)
where f (t) is a known function of time.

Figure 2.9: (a) Symbol of pure source of velocity. (b) Actuator of prescribed relative velocity and (c) its
multipole model.

Fig. 2.9a gives our graphical symbol for the pure source of velocity. The two pins sticking out of
the circle denote the source poles representing the energy interaction of the source with the rest of the
system. Besides the prescribed velocity v(t), the source is associated with its force F (t). This force
counterbalances the forces imposed on the source by the system in which the source operates. The
source polarity denotes the + sign by one of its poles. The assumed positive orientation of v(t) and
F (t) in relation to the symbol polarity is indicated in Fig. 2.9a by the empty-head and full-head arrow,
respectively.
Fig. 2.9b shows a real rectilinear actuator of negligible mass acting between the interaction sites A
and B so that ẋBA (t) = va (t) where va (t) is the actuator velocity. Let us assume that va (t) is sufficiently
independent from the external forces Fext and −Fext exerted on the actuator by the system. Then the
actuator can be modelled by a pure source of velocity as shown in Fig. 2.9c. The rectilinear motion of
the interaction sites A and B in the x-direction is represented in the model by nodes Ax and Bx .
Fa and −Fa are the internal actuator forces counterbalancing the external forces. Let us assume that
the source symbol modeling the actuator is placed in the diagram in such a way that the orientation of
its velocity v is in agreement with the velocity va of the actuator. Then the source force F represents the
internal actuator force Fa acting at the site of interaction represented by the node to which the + pole
of the source is attached.
2.5. PRESCRIBED RECTILINEAR INTERACTIONS 2-9

The power consumed by the source of velocity is


P(t) = v(t) · F (t) (2.22)
If the velocity v and force F of the source are either both positive or both negative the modelled actuator
consumes energy supplied by the remaining system. In the former case, the system forces the actuator to
extend its length whereas in the latter case the actuator length is shortened by the system. If, however,
the signs of the source velocity v and force F are different the modelled actuator delivers energy into the
system. If both signs are positive the actuator pushes the system sites apart, if both signs are negative
the sites pulled closer to each other.

Example:

Figure 2.10: (a) Car towing, (b) its model. (c) Towing velocity profile, (d) corresponding force exerted
by towing car, (e) energy accumulated in towed car, (f ) position of both cars.

The most trivial dynamic model of the two-car system shown in Fig. 2.10a is given in Fig. 2.10b. As
it is assumed that all the system bodies are rigid and the towing rod is massless a single point – the point
A – is sufficient for representing the system rectilinear motion along x. Only inertia is considered in the
towed car dynamics and the towing car is modelled by a source of velocity.
Let us consider the case when the velocity profile of the towing car velocity vc2 is triangular (Fig. 2.10c).
The source force Fc2 response (Fig. 2.10d) is then negative constant when the velocity increases, and
positive constant when the velocity decreases. Therefore, the towing car supplies energy into the system
in the former case whereas it consumes energy in the latter case. In both cases, the energy accumulated
in the towed car is positive (Fig. 2.10e).
The forces Fc1 and Fc2 are in balance, i.e.
Fc1 (t) + Fc2 (t) = 0 (2.23)
In conformity with our conventions, the arrows indicating orientation of the forces Fc1 and Fc2 oppose the
x-direction in the system geometric representation, while they are directed away from the non-reference
node Ax in the model of the system.

2.5.2 Indicators of position and acceleration


Unlike velocities or forces, positions and accelerations are not primary variables representing rectilinear
motion in multipole diagrams. Nevertheless, they can be indicated very easily using blocks: integrators
for positions and differentiators for accelerations.
To demonstrate this, Fig. 2.11a shows indication of the car position xA in the car-towing model
given in Fig. 2.10b. The input of an integrator is attached to the node Ax . Thus the velocity ẋA (t)
associated with this node is integrated to obtain at the integrator output the position xA (t) of the point
A representing the car motion: Z t
xA (t) = v(τ )dτ + xA (t0 )
t0
2-10 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

Figure 2.11: Indication of (a) position and (b) acceleration.

The initial condition of the integration xA (t0 ) corresponds to the initial position of point A at t = t0 .
The car position is for the triangular velocity profile shown in Fig. 2.10f .
In a similar way, a differentiating block can be used to indicate the car acceleration. This illustrates
Fig. 2.11b in which the differentiator is attached to the node Ax . The indicated acceleration

d
ẍA = ẋA
dt
is denoted at the differentiating block output as dAx.
Note that attaching a block input to a node does not disturb the force balance at the node as no force
enters any block.

2.5.3 Prescribed forces


To model actuators of prescribed forces we need to introduce a pure source of force. In general, such
a source exhibits a force F (t) characterized by the constitutive relation

F = f (t) (2.24)

where f (t) is a known function of time.


Our graphical symbol for the source of force is given in Fig. 2.12a. The full-head and empty-head
arrows indicate the assumed positive orientation of the force F (t) and velocity v(t) associated with the
source. The variable orientation is related to polarity of the source symbol denoted by the + sign. The
relative velocity v(t) of the source represents the velocity imposed on the force source by the rest of the
system in which the source operates.

Figure 2.12: (a) Symbol of pure source of force. (b) Actuator of prescribed force and (c) its multipole
model.

Fig. 2.12b shows a real rectilinear actuator of a prescribed force Fa acting along the x-axis between
the interaction sites A and B of a system. The actuator force Fa is counterbalancing the external force
Fext to the actuator. The effect of the real actuator is modelled in Fig. 2.12c using the pure source of
force. The motion of the interaction sites A and B in the x-direction is represented in the model by the
nodes Ax and Bx .
Polarity of the full-head arrow for the force Fa indicates that if Fa > 0 the actuator pulls the sites A
and B closer. If Fa < 0 the actuator pushes the sites A and B apart.
The source symbol should be oriented between the nodes in agreement with the orientation of the
x axis associated with the real actuator. Therefore, the + pin of the symbol should be attached to the
node Bx representing the actuator site of interaction B that is assumed to move ‘faster’ then its site A. If
2.5. PRESCRIBED RECTILINEAR INTERACTIONS 2-11

the source is oriented properly its force F represents the prescribed force Fp acting at the site B related
to the + pole of the source.
If the source symbol modeling the actuator is placed in the diagram in such a way that the orientation
of its velocity v is in agreement with the velocity va of the actuator, the source force F represents the
actuator force Fa at the actuator end which is related to the + pole of the source.
If the source velocity v and force F are either both positive or both negative the modeled actuator
consumes energy supplied by the remaining system. In the former case, the system forces the actuator to
extend its length whereas in the latter case the actuator length is shortened by the system. If, however,
the signs of the source velocity v and force F are different the modeled actuator delivers energy into the
system. If both signs are positive the actuator pushes the system sites apart, if both signs are negative
the sites pulled closer to each other.

Examples:

Figure 2.13: Car towing.

The free vertical fall of a ball shown in Fig. 2.14a is modeled in Fig. 2.14b using a source of force Fg
to model the gravitational force Fg = 9.81m acting on the body. The ball air resistance is neglected in
the model.

Figure 2.14: (a) Free fall of ball, (b) its model.

2.5.4 Indicators of velocities and forces


Pure sources of zero force can be used as physical elements modeling ideal indicators of velocity, i.e.
instruments measuring relative velocity of distinct sites of interaction. Our graphical symbol for such
an indicator is given in Fig. 2.15a. Direction of the arrow showing the assumed positive polarity of the
indicated velocity is fixed to the polarity of the symbol.
Fig. 2.15b shows this symbol utilized to indicate the relative velocity of a force actuator modelled by
the pure source of force placed between nodes Ax and Bx .

Figure 2.15: (a) Symbol of ideal indicator of velocity. (b) Indication of a source-of-force velocity.

Pure sources of zero velocity can be utilized as physical elements modeling ideal measuring instruments
or indicators of force. Our graphical symbol for such an indicator is given in Fig. 2.16a. Direction
of the arrow showing the assumed positive polarity of the indicated force is fixed to the polarity of the
2-12 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

symbol. Fig. 2.16b shows this symbol utilized to indicate the force of a source of velocity representing
the force exerted on a velocity actuator by the system in which it operates.
Note the difference between measuring velocity and force. The direct measurement of forces requires
disconnecting the system and connecting it back by a force indicator. Velocities (as well as positions
or accelerations) are measured by indicators applied to systems without any need to disconnect them.
This is the reason why velocities represent rectilinear across variables while forces are rectilinear
through variables.

Figure 2.16: (a) Symbol of ideal indicator of force. (b) Indication of a source-of-velocity force.

Re-solvable examples
Body driven by given force
Body driven by given velocity
Vertical throw of a ball
Body driven by a frequency-variable force

2.6 Rectilinear damping


Almost every mechanical motion is associated with some inherent degree of friction which resists the
motion. Sometimes friction is unwanted but must be tolerated and accounted for, as, for example, in
bearings or in the aerodynamic drag on a moving body. In other cases friction is desired and is designed
into equipment, as in brakes, clutches, belt drives, and wedges. Wheeled vehicles depend on friction for
both starting and stopping, and ordinary walking depends on friction between the shoe and the ground.
Friction is utilized in vibration dampers and shock absorbers such as dashpots used in car suspension
systems to damp car oscillations caused by riding over bumps in the road.

Figure 2.17: (a) Symbol of pure damper, (b) F − v characteristic of the ideal damper.

Fig. 2.17a gives our graphical symbol for pure rectilinear dampers. A pure rectilinear damper
exerts a resistive force F that is a function of the relative velocity v of its motion, but exhibits no mass
or spring effects. Thus,
F = f (v) (2.25)
where f (·) is a function such that F = 0 when v = 0. In the case of an ideal rectilinear damper this
constitutive relation is linear, i.e.,
F (t) = b · ẋ(t) (2.26)
where b [N.s/m] is the damping constant of the ideal damper. The ideal damper characteristic is shown
in Fig. 2.17b.
Characteristics close to those of an ideal damper demonstrate, for example, the sliding surfaces shown
in Fig. 2.18a if they are separated by a thin layer of a fluid lubricant, or the dashpot in Fig. 2.18c if it
does not move too fast. In such cases the friction forces result from the viscous drag of a fluid under
laminar flow conditions.
2.6. RECTILINEAR DAMPING 2-13

Fig. 2.18a shows sliding surfaces in contact pushed towards each other by some forces Fn normal to
the motion. The frictional forces Fr resisting a relative motion of the surfaces must be compensated for
by external forces Fext .

Figure 2.18: (a) Sliding surfaces in contact, (b) corresponding model. (c) Oil-filled shock absorber.

In Fig. 2.18b, the friction effect between sliding surfaces is modelled by an ideal damper. Nodes Ax
and Bx in the model represent x-motions of points A and B of the real surfaces in the x-direction. If the
surfaces are well lubricated by a fluid, the damping constant


b= (2.27)
d
where A is the common area and d is the thickness of the lubrication film between the surfaces, ν is the
viscosity coefficient of the lubricant.
Note the way in which the polarities of the damper relative velocity v and resistive force F are
associated with the damper symbol shown in Fig. 2.17a. As the symbol is asymmetric, no + sign is
necessary to denote the symbol polarity (as it was the case of pure sources). If the damper symbol is
placed in a multipole diagram in such a way that its velocity v represents polarity of the relative velocity
ẋBA , then the damper force F represents the resistive force Fr at the ‘faster’ site of interaction. The
damper force F thus corresponds in Fig. 2.18b to the resistive force counterbalancing the net external
force Fext acting on point B in the real bodies.
Since no mass is associated with the surfaces, the external forces Fext applied to the interaction sites
A and B are balanced, and the same must be true about the reaction forces Fr at these sites. Thus an
external force applied to point B is assumed to flow through the damper to point A without any change.
The shock absorber shown in Fig. 2.18c consists of a piston inside an oil-filled cylinder in which fluid
is forced through a narrow orifice in the piston by relative motion, thus producing resisting force. In this
case, the damping constant can be approximated as
" 2 #2
D
b = 8πνℓ −1 (2.28)
d

where D is the cylinder diameter, d is the diameter of the orifice, and ℓ is the length of the orifice.
The mechanical power dissipated in the pure damper as heat is

P = ẋ · F

Note that it is always positive regardless of the direction of the damper relative motion.

Re-solvable examples
Viscous friction experiment
Two-mass model of car
2-14 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

2.7 Rectilinear compliance


2.7.1 Rectilinear deformations
Materials undergo a deformation when they are compressed or stretched. If a mechanical component
experiences a steady deformation when acted on by steady forces, it is exhibiting compliance. The
component compliance can be modelled by pure springs. As these physical elements ignore any inertial
and dissipative effects, their deformation is a function only of the applied forces. Fig. 2.19a gives our
symbol of pure springs. The symbol resembles a wire coil spring shown in Fig. 2.19b, the most familiar
mechanical device design to exhibit compliance.

Figure 2.19: (a) Symbol of pure spring. (b) Coil spring stretched and compressed. (c) Ideal-spring
deformation characteristic.

The spring deformations due to compliance are often approximated to be proportional to the ap-
plied force. A linear pure spring, which is called an ideal spring, has a proportional relation between
deformation and force,
∆ℓ = L · F (2.29)
where ∆ℓ [m] is the spring deformation, F [N] is the spring restoring force, and the constant parameter L
[m.N−1 ] is the compliance factor of the spring. The reciprocal value of the compliance factor, K = 1/L
[N.m−1 ], is the spring stiffness termed also the spring constant.
For example, the compliance factor of a coil spring can be derived using the methods of applied
mechanics to be approximately
8D3 n
L=
Gd4
where D [m] is the spring pitch diameter (mean coil diameter), d [m] is the diameter of the wire, n [–] is
the number of coil turns, and G [N.m−2 ] is the shear modulus of the coil material.
The compliance factor of a cylindrical bar made of homogenous ideally elastic material is according
to Hook’s law
ℓ0
L= (2.30)
E ·A
where ℓ0 [m] is the force-free length of the rod, A [m2 ] is the rod cross-sectional area, and E [N.m−2 ] is
the Young’s modulus of elasticity of the rod material.
Fig. 2.19b shows a stretched, force-free and compressed spring. The spring deformation ∆ℓ = ℓ − ℓ0
where ℓ is the spring actual length ℓ = xB − xA and ℓ0 is the free length of the spring. If ∆ℓ > 0 the
spring is stretched, if ∆ℓ < 0 the spring is compressed. Fig. 2.19c shows the ℓ − F characteristic of the
ideal spring.
By differentiating (2.29), we can arrive at the constitutive relation of an ideal rectilinear springs

d
v =L· F (2.31)
dt
where v = dℓ/dt is the spring relative velocity. Alternatively, the relation (2.31) can be expressed in the
integral form
Z t
F =K v dτ + F (t0 ) (2.32)
t0

Fig. 2.20a shows a coil spring placed between the interaction sites A and B in the x-motion. If the
spring mass is negligible in comparison with masses of the interacting components it can assumed that
2.7. RECTILINEAR COMPLIANCE 2-15

Figure 2.20: (a) Coil spring, (b) force-controlled and (c) elongation-controlled model. (d) Model respecting
spring mass and damping. (e) Distributed-parameter model.

the external forces Fext applied to the spring end-points are balanced. This implies that also the restoring
forces Fr , opposing the external forces, are balanced at the spring end-points.
Note that one of the pins of the pure-spring symbol given in Fig. 2.19b is denoted by the + sign.
This allows for an unambiguous association of the polarities of the spring velocity v and the spring force
F with the polarity of the symbol. In Fig. 2.20b, the pure-spring symbol is oriented in conformity with
the chosen orientation of the modelled real spring shown in Fig. 2.20a. Then the pure-spring force F
represents the real-spring restoring force Fr at the assumed ‘faster’ end of the real spring.
To express the total length of the modelled spring, the pure spring is combined with an integrator in
Fig. 2.20b. For a chosen pre-loading force F (t0 ) of the spring, the initial relative distance between the
spring end-points at t = t0 should be specified as

xBA (t0 ) = ℓ0 + L · F (t0 ) (2.33)

where ℓ(t0 ) is the initial length of the spring.


Fig. 2.20c gives an alternative elongation-controlled model to the ideal rectilinear spring shown in
Fig. 2.20b. The ideal spring is replaced there by a source of the restoring force

Fr = k [xBA (t) − ℓ0 ] (2.34)

The source is controlled by the relative distance of the spring end points xBA . The initial value of this
distance for a chosen initial length of the spring ℓ(t0 ) is

xBA (t0 ) = ℓ(t0 ) (2.35)

Figure 2.21: (a) Suspended ball, (b) its dynamic model, (c) ball position.

Examples:
Fig. 2.21b shows a dynamic model of the ball suspended on a spring given in Fig. 2.21a. Node Ay
represents the y-motion of point A of the ball. The spring symbol is placed between the node Ay and
the reference node in such a way that the polarity of its velocity v is in agreement with the polarity of
the relative velocity of the real spring along the y axis. As a result, the pure spring force F represents,
2-16 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS

both by its magnitude and polarity, the actual restoring force counterbalancing the gravitational force Fg
acting on point A in the real bodies. The effect of the gravitational force Fg = 9.81m is modelled using
a source of force.

Figure 2.22: (a) Car following bumpy road, (b) car-wheel suspension, (c) quarter-car model.

Fig. 2.22a shows a car following a bumpy road while Fig. 2.22b shows just a detail of the car, the
suspension of one of its wheels. In Fig. 2.22c point Ry represents the y-motion of the contact point
between the car tyre and the road surface R when the car moves in the x-direction. Points Wy and By
represent the y-motion of the wheel axis W and the body-suspension interaction point B, respectively.
Lt and bt is the compliance and damping of the wheel tyre, respectively. Similarly, Ls and bs is the
compliance and damping of the wheel tyre, mw is the mass of the complete wheel and mb is the quarter
of the mass of the car body. The model is excited by the source of the velocity ẏR resulting from the tyre
copying the road bumps.

2.7.2 Spring potential energy


From (2.29) and (2.30) we can see that the energy flow into an ideal rectilinear spring at a given instant
is
P = ẋ · F = L · F · Ḟ (2.36)
If P > 0, energy is being stored in the spring, and if P < 0, the stored energy is being retrieved from the
spring. Which of these two states is occurring depends both on the spring deflection and relative velocity
at the moment.
Integration of (2.36) gives an expression for the energy accumulated in the ideal rectilinear spring
1
E= L · F2
2
which is called rectilinear potential energy.
We note that this energy depends directly on the force transmitted through the spring, but not on the
velocity difference. Hence the spring stores energy by virtue of its force. This energy is always positive
and does not depend on the sign of either the force or the spring deflection.

2.8 Summary
To set up a dynamic model of a real rectilinear system and to simulate its behavior using DYNAST, you
should proceed through the following steps.
2.8. SUMMARY 2-17

1. Sketch the geometric configuration of your system, i.e.:


• clearly separate the system of your concern from the system surroundings and determine a
reference frame for the system
• associate the rectilinear motions with a coordinate axis
• for each body denote the geometric point representing its rectilinear motion and give it a name
• assume the positive velocity of each such point in the direction of the chosen coordinate axis
2. Set up a multipole diagram representing the dynamic model of your system and its surroundings
(you may use directly the DYNAST schematic editor):
• for each body place a node into the diagram corresponding to the geometric point representing
the body velocity
• give each node a name identical with the name of the point (using a label in the DYNAST
schematic editor)
• associate each of the nodes with the symbol for inertors
• place into the diagram one or more symbols for reference nodes representing the reference
frame
• place symbols for elements representing couplings between the nodes (if necessary, interconnect
the element pins with the nodes by links)
• add symbols for elements representing the system surroundings
• invert the element symbols if their velocity is oriented in the opposite direction with respect
to the chosen coordinate axis
• give names to the inertors and the other elements and specify values of their parameters
3. Specify analysis of your system model:
• select the Numerical Nonlinear Analysis
• specify the required Time Interval of the analysis and set the values of Initial Conditions
• select Desired Variables like node and element velocities, element forces, and block outputs
• To invoke the analysis, select Run Analysis & Plot in menu Run
4. Interpret the analyzed results:
• Set the desired variables in menu Plot
• in menu Axes choose the way you want to plot the variables
• the assumed positive polarity of velocities are oriented in the direction of the chosen coordinate
axes whereas the assumed positive polarity of forces of the elements is opposite to this direction
2-18 MODULE 2. RECTILINEAR MECHANICAL SYSTEMS
Module 3

Rotational mechanical systems

3.1 Rotational interactions


3.1.1 Rotational motion
In rotation about a fixed or nonaccelerating axis all points of the site other than those on the axis move
in concentric circles about the fixed axis.

Figure 3.1: Subsystems in rotational motion.

Fig. 3.1 shows a subsystem with two concentric shafts rotating about the z-axis fixed to the reference
frame. The subsystem might be a torsional spring, a rotational damper, or a gearbox, for example. The
ends of the shafts A and B are in contact with shafts of some other subsystems. The power intake of the
subsystem due to the rotational motion of its energy entries A and B about the z axis is

P(t) = ϕ̇A (t) · τA (t) + ϕ̇B (t) · τB (t)

where ϕ̇A (t) and ϕ̇B (t) are absolute angular velocities of the entries A and B about the z axis. τA (t) and
τB (t) internal torques counterbalancing external torques τA ext(t) and τB ext(t) applied to the entries.
Absolute angular motion about the axis z is described in Fig. 3.1 by the angular displacement
ϕ, i.e., by the angle subtended between a line drawn on the rotating entry which lies in the x-y-plane
perpendicular to the axis of rotation z and the coordinate axis x fixed to the reference frame. In this
text, the counterclockwise direction will be considered as the positive orientation of rotation in a plane
perpendicular to the axis of rotation.
Absolute angular velocity ϕ̇(t) [rad/s] and absolute angular acceleration ϕ̈(t) [rad/s2] of an
entry is by definition
dϕ dϕ̇ d2ϕ
ϕ̇(t) = and ϕ̈(t) = =
dt dt dt2
where ϕ(t) is the angular displacement of the entry. The relative angular displacement, relative
angular velocity and relative angular acceleration of the entry B with respect to the entry A is

ϕBA = ϕB − ϕA ϕ̇BA = ϕ̇B − ϕ̇A ϕ̈BA = ϕ̈B − ϕ̈A

3-1
3-2 MODULE 3. ROTATIONAL MECHANICAL SYSTEMS

3.1.2 Torque
A moment of force or torque τ [N.m] can be viewed as the time rate of the angular momentum
ρ(t) acting in the same direction, i.e.

τ=
dt
The complete specification of a moment of force τ (t) must include its magnitude and the axis of rotation.
To have the energy flow through an entry into a subsystem positive, the torques acting on the entry must
be oriented in the opposite way with respect to the direction of the entry angular velocity.

3.2 Ideal rotational inertors


An ideal rotational inertor models the rotational motion of a rigid body about an axis fixed to the
reference frame. The constituency relation of this physical element corresponds to Newton’s second law
which can be written as
ρ(t) = J ϕ̇(t) (3.1)
or as
d
τ (t) = J ϕ̇ (3.2)
dt
where ρ(t) is the element rotational momentum, τ (t) is the element moment of force, ϕ̇(t) is the element
angular velocity, and J [kg.m2 /rad] is the moment of inertia of the element.

Figure 3.2: (a) Symbol of rotational inertor. (b) Flywheel, and (c) its model.

Rotational inertors model inertial effect in rotational motion of bodies exclusively, they ignore any
damping or spring effects as shown in Fig. 3.2. The angular velocity ϕ̇(t) of a rotational inertor represents
the absolute angular velocity of a body measured in relation to a nonaccelerating reference frame. Thus
the −pole of rotational inertors must be always coalesced with a node corresponding to the reference.
This is the reason why only the +pole is shown in the element symbol.
If the moment of inertia JT of a body about a fixed axis passing through the mass center T is known,
the moment of inertia J of the body about another fixed axis parallel to the former one can be determined
as
J = JT + md2 (3.3)
where m is the mass of the body, and d is the separation of the axes.This formula, known as the parallel-
axis or Steiner’s theorem, is illustrated by Fig. 3.3. The diameter of the ball of the mathematical
pendulum of Fig. 3.3a is so small that its moment of inertia is negligible. If also the mass of the pendulum
string is neglected, the pendulum moment of inertia with respect to the axis of rotation A is J = mr2 .
Fig. 3.3b shows a compound pendulum. If only the moment of inertia JT with respect to mass center
of the compound pendulum body is known, the pendulum moment of inertia with respect to the axis of
rotation A must be determined using (3.3). The mass center of each of the pendulums is acted on by the
moment of the gravitational force Fg = mg, the moment arm of length r sin ϕ varies with the pendulum
deflection ϕ. In the dynamic model for both pendulums shown in Fig. 3.3c, the ideal rotational damper
of factor B respects damping of the pendulums.

3.3 Ideal rotational dampers


The constituency relation of an ideal rotational damper with the symbol of Fig. 3.4a is
τ (t) = br · ϕ̇(t) (3.4)
3.3. IDEAL ROTATIONAL DAMPERS 3-3

Figure 3.3: (a) Mathematical and (b) compound pendulum, (c) pendulum dynamic model.

where τ (t) is the moment of force, ϕ̇(t) is the relative angular velocity, and br [N.m.s/rad] is the rotational
damping constant of the damper.

Figure 3.4: (a) Symbol of pure rotational damper. (b) Oil-filled rotational damper, and (c) its model.

An ideal rotational damper is a twopole model that ignores any mass or spring effects and assumes
that a moment of force is proportional to the relative angular velocity of its poles. An example of a real
device with the τ -ϕ̇ characteristic close to that of the ideal damper, is the viscous rotational damper
shown in Fig. 3.4b. It is a propeller rotating in oil inside a cylinder and producing the resisting moment
of force.
The poles represent ϕ-motions of the energy entries B and A of a real damping subsystem. τ represents
the internal moment of force at the entry associated with the +pole, i.e. at the entry B, counterbalancing
an external moment of force applied to this entry. The positive direction of τ is opposite to that for the
damper angular velocity ϕ̇ corresponding to the relative velocity of the entries ϕ̇BA = ϕ̇B − ϕ̇A .
Since an ideal rotational damper has no moment of inertia, the moments of force acting on it must be
balanced. Thus, an external moment of force applied to Bϕ in the ϕ-direction flows through the damper
and out through Aϕ without a change and, at the same time, the internal moment of force at the damper
−pole denoted by Aϕ must be −τ . Note that the power flow into the ideal damper P = ϕ̇τ is always
positive regardless of the direction of its relative motion.
Fig. 3.5a shows the symbol of the ideal rotational damper used to model a shaft passing through a
well lubricated journal bearing fixed to the reference frame. In Fig. 3.5b, the same model is utilized for
an oil-polluted clutch the discs of which are slipping with respect to each other so that they rotate with
different angular velocity and there is a viscous friction effect between them.

Figure 3.5: Ideal model of (a) a shaft in journal bearing, and (b) a slipping clutch.
3-4 MODULE 3. ROTATIONAL MECHANICAL SYSTEMS

3.4 Ideal rotational springs


For an ideal rotational or torsional spring, the constituency relation is

ϕ(t) = Lr τ (t) (3.5)

or
d
ϕ̇(t) = Lr τ (t) (3.6)
dt
where ϕ(t) is the ideal spring angular displacement, ϕ̇(t) is the ideal spring angular velocity, τ (t) is
the ideal spring moment of force, and Lr [rad.N−1 .m−1 ] is the torsional compliance of the spring. The
reciprocal value of Lr is the spring torsional stiffness or torsional spring constant.

Figure 3.6: (a) Symbol of pure torsional spring. (b) Spiral torsional spring, and (c) its model. (d)
Helical-coil spring. (e) Twisted rod.

The pure spring has no inertia by definition and hence transmits torque undiminished. One of the most
common mechanical components which exhibit the characteristics of a rotational spring is the torsional
spiral spring shown in Fig. 3.6b. Fig. 3.6c shows a helical-coil torsional spring, and Fig. 3.6d a twisted
rod. ϕBA corresponds to the relative angular deformation of the real spring, while τ represents the spring
internal moment of the force counterbalancing the external moment of a force applied to the spring.

Figure 3.7: (a), (b) torsional pendulum. (c), (d) double torsional pendulum.
Module 4

Coupled mechanical systems

4.1 Translatory-to-translatory couplings


4.1.1 Translational transformer
Translatory-to-translatory coupling is the coupling between the translational motion of one point
and the translational motion of another point along the same or different axis. Such couplings can be
easily modelled by the pure translational transformer. It is an idealized mechanical device which
neither dissipates nor stores any energy and determines ratios of velocities and forces associated with
the two coupled motions. Fig. 4.1 shows our symbol for pure transformers. The variables associated in
Fig. 4.1 with this symbol correspond to the special case when the symbol is applied to a translational
transformer.

Figure 4.1: Pure translational transformer.

Motion of point A with respect to point A along the x axis is represented by poles Ax and A x ,
respectively. Similarly, motion of point B with respect to point B along the ξ axis is represented by
poles Aξ and A ξ . Fx is the force exerted by the transformer to move points A and A apart while
counterbalancing external forces applied to these points. Fξ is the force counterbalancing external forces
applied to points B and B .
The relation between velocities associated with the translational transformer is defined as

ẋ/ξ˙ = n (4.1)

where ẋ is the relative velocity of point A with respect to point A , and ξ̇ is the relative velocity of point
B with respect to point B , n is the transformer ratio. An ideal transformer has this ratio constant.
Forces associated with the translational transformer are related by definition as

Fξ /Fx = −n (4.2)

The power put into the translational transformer is

P = Fx ẋ + Fξ ξ˙ (4.3)

After submitting (4.1) and (4.2), (4.3) becomes P = 0. A pure translational transformer thus transforms
a translational motion into another translational motion and vice versa. It performs this function with
no energy loss or storage, so that the net power flow into a pure transformer is zero.

4-1
4-2 MODULE 4. COUPLED MECHANICAL SYSTEMS

Table 4.1: Examples of translatory-to-translatory transducers.

Transducer Coupling ratio n

a
b

a

b

−1

ra

rb

Table ?? shows examples of pure models of some translatory-to-translatory transducers. They include
levers, pulleys and a rigid link of a fixed length the endpoints of which are constrained to a linear or
circular path.

4.1.2 Levers
A translatory-to-translatory coupling is typified by a lever which has no mass, is infinitely stiff, and is
free from frictional effects. Such a lever is shown in Fig. 4.2a.

Figure 4.2: (a) Lever, (b) its dynamic model.

The relationship between the velocities of endpoints A and B of the lever with respect to the pivot C
is
ẏA /ẏB = −a/b = n (4.4)
where n is the lever ratio. Note that a and b change as ϕ varies, but that a/b remains constant. Often
ϕ will be limited to a small angle, so that a and b are essentially constant.
The full-head arrows denote forces FA , FB and FC exerted by the lever at its end points A and B and
at the pivot C, respectively, to counterbalance external forces applied to these points. Since the massless
lever has no momentum, then by definition the sum of the forces in any direction and the moments of
the forces about the pivot must be zero:
FA + FB + FC = 0 (4.5)
4.2. ROTARY-TO-ROTARY COUPLINGS 4-3

FA a + FB b = 0 (4.6)

Therefore,
FB /FA = a/b = −1/n (4.7)
Note, that the relationships between velocities (4.4) and between forces (4.7) are the same as in the
translational transformer used in Fig. 4.2b to model the lever.

4.2 Rotary-to-rotary couplings


4.2.1 Rotational transformer
Rotary-to-rotary coupling is the coupling between the rotational motion about a fixed axis and
another rotational motion about the same or different axis. Such couplings can be easily modelled by
the pure rotational transformer. It is an idealized mechanical device which determines ratios of
angular velocities and torques associated with the coupled motions. Fig. 4.3 shows our symbol for pure
transformers. The variables associated with this symbol in Fig. 4.3 correspond to the special case when
the symbol is applied to a rotational transformer.

Figure 4.3: Pure rotational transformer.

Rotational motion of point A with respect to point A rotating about the same axis with angular
velocity ϕ̇ is represented by poles Aϕ and A ϕ , respectively. Similarly, the rotational motion of point
B with respect to rotating point B with the velocity ψ̇ is represented by poles Aψ and A ψ . τϕ is the
torque exerted by the transformer to move points A and A apart while counterbalancing external torques
applied to these points. τψ is the force counterbalancing external torques applied to points B and B .
Relation between angular velocities associated with the rotational transformer is

ϕ̇/ψ̇ = n (4.8)

where ϕ̇ is the relative angular velocity of point A with respect to point A , and ψ̇ is the relative angular
velocity of point B with respect to point B . n is the transformer ratio. An ideal transformer has this
ratio constant.
Torques associated with the rotational transformer are related as

τψ /τϕ = −n (4.9)

The power put into the rotational transformer is

P = τϕ ϕ̇ + τψ ψ̇ (4.10)

After submitting (4.8) and (4.9), (4.10) becomes P = 0.


A pure rotational transformer is thus an idealized mechanical device which transforms a rotational
motion into another rotational motion and vice versa. A pure transformer performs this function with
no energy loss or storage, so that the net power flow into a pure transformer is zero.

4.2.2 Gear trains


The simplest and most familiar example of a rotary-to-rotary coupling is the elementary gear train shown
in Fig. 4.4a. Two mating gears of different sizes are attached to two parallel shafts which in turn are free
to rotate supported in a housing. Gear-tooth profiles machined into the periphery of the gear disks cause
the gears to behave like circular cylinders of radii ra and rb , which roll on each other without slipping,
4-4 MODULE 4. COUPLED MECHANICAL SYSTEMS

provided that the gears are rigid and the gear teeth are geometrically perfect. The radii ra and rb are
called the pitch radii of the gears, and their magnitudes are proportional to the numbers of teeth Na and
Nb . The gear ratio is defined as
ra /rb = Na /Nb

tA tA jA
jA tA tB

tB
tB jA jB
jB
jB
(a) (b) (c)

Figure 4.4: (a) Gear train, (b) its free-body diagram, (c) transformer model.

Since the gears are assumed to roll without slipping, the arc distance ra ϕ̇A ∆t turned off on gear a
must equal minus the arc distance rb ϕ̇B ∆t turned off on gear b. Hence
ϕ̇A rb Nb
=− =− =n (4.11)
ϕ̇B ra Na
where the negative sign indicates that the direction of ϕ̇B is opposite to that of ϕ̇A .
To determine the relationship between the torques τA and τB , let us isolate the two gears from each
other and from the frame, thus forming their free-body diagrams of Fig. 4.4b. Only torques and torque-
producing forces are shown on these free bodies, and friction between the gear-tooth surfaces is assumed
to be zero. If we further assume that the gears have zero inertia or constant speed, the net torque on
each gear is zero. If we also assume the bearing torques to be negligible,
τA = Ft ra and τB = Ft rb
Employing now the definition of n in (4.11), we obtain
τA ra Na
= = = −1/n (4.12)
τB rb Nb
Thus the speed is changed as n while the torque is changed as 1/n. The net power into the gear train
for ϕ̇A = 0 is therefore zero:
P = τA ϕ̇A + τB ϕ̇B = τA ϕ̇A + (τA /n)(−τA n) = 0
Since the gear system is assumed to have constant angular momentum, the torque τC shown acting
on the container Fig. 4.4a must be equal to minus the sum of τA and τB , or
τA + τB + τC = 0 (4.13)
Since ϕ̇C = 0 in the preceding discussion, τC does no work on the system.
Table 4.2 shows various setups of gear trains that can be modeled by a single pure rotational trans-
former. The table gives the corresponding gear ratios.

4.2.3 Gear backlash


Fig. 4.5 indicates the way in which a pure model of a pair of spur gears can be augmented to take into
the consideration backlash between the gears. Fig. 4.5b shows a simple characteristic approximating the
torque τ exerted by the gears on each other in relation to the angle ϕAB between the gears. The finite
slope of the nonzero branches of the characteristic corresponds to the torsional stiffness of the gear teeth.
The backlash effect is implemented in the model by adding the source of torque τ controlled by the angle
ϕAB in accordance with the characteristic shown in Fig. 4.5b to the energy-transfer transformer modeling
the backslashes motion conversion. Also slippage in rotary-to-rotary motion converter due to friction can
be modelled in a similar way.
4.2. ROTARY-TO-ROTARY COUPLINGS 4-5

Table 4.2: Rotary-to-rotary motion converters

Gear train Configuration Coupling ratio n

External rb
spur gears −
ra

Internal rb
spur gears
ra

Bevel rb
gear pair
ra

Planet rb
gear
ra

Skew rb
gear pair
ra
4-6 MODULE 4. COUPLED MECHANICAL SYSTEMS

tB tA tB
jB

tA jA
jAB jAB jA t(jAB) j
B

(a) (b) (c)

Figure 4.5: (a) Gear train with backslash, (b) backslash characteristic, (c) multipole model.

4.2.4 Belt and chain systems

tB tB
jA
tA
jA

jB jB
tA
(a) (b)

Figure 4.6: Belt-and-pulley and chain-and-sprocket assemblies.

Fig. 4.6 shows pure belt-and-pulley and chain-and-sprocket systems that perform essentially the same
function as gear trains. As opposed to a belt and pulley, the latter system does not allow slippage between
the belt and sprocket due to the sprocket teeth. Note that the sign of n is positive in the case of the
system in Fig. 4.6a as the pulleys rotate in the same direction, whereas in the case of the system in
Fig. 4.6b the sign is negative.

4.2.5 Belt friction


Fig. 4.7a shows a flat belt passing over a fixed cylindrical drum. To model belt slipping let us consider
interactions between segments of the belt and drum surface which are in contact. This can be taken
as a one-dimensional translational problem if the belt segment as well as the adhering drum surface is
straighten up as shown in Fig. 4.7b. The variables can be then related to the coordinate axis x.

Figure 4.7: Belt friction.

The Eytelwein’s formula gives the relation between the forces FA and FB acting at the belt-segment
endpoints A and B. For the case when the belt is just about to slip if pulled to the right, i.e. if FA > FB > 0,
4.2. ROTARY-TO-ROTARY COUPLINGS 4-7

the formula is:


FA
= eµγ (4.14)
FB
where µ [–] is the coefficient of friction, and γ [rad] is the angle of the belt contact. This formula is
derived under the assumption that the belt segment is massless and rigid, and the friction between the
belt and the drum is governed by the ‘law of dry friction’.
A multipole diagram of the belt friction constructed under the above assumptions is given in Fig. 4.7c.
The contact areas of the belt and drum segments are represented there by poles A and C, respectively.
Friction is modelled by a source of friction force Ff acting between these two poles. The rigid belt
segment is modelled by the source of zero velocity ẋAB placed between the poles A and B representing
the belt-segment end points. At the same time, this source is employed in the model as a sensor of the
tension in the belt segment. This tension force Ft is used to control the friction-force source Ff .
Taking into consideration the balance relations for forces at the poles
FA = Ft − Ff and FB = −Ft (4.15)
and substituting them into the formula we shall obtain the constitutive relation of the friction-force source
Ff = Ft 1 + eµγ·sgn(Ft )
 
(4.16)

The function sgn(Ft ) was included into the formula to make it valid also for the case when the belt
is pulled to the left, i.e. for FB > FA > 0. (You are encouraged to check its correctness.)
Equation (4.16) can be used both for static friction of impending slipping as well as for the slipping
kinetic friction after substituting there either the coefficient of static friction µs or the coefficient of kinetic
friction µk for µ. Fig. 4.8 shows an example of the Ff -ẋ characteristic of combined static and kinetic
friction governed by (4.16) for the tension Ft = ±1. Note that the characteristic is asymmetric with
respect to the origin.

Figure 4.8: Belt friction characteristic.

As shown in Fig. 4.7d, the belt friction model from Fig. 4.7c can be completed to respect the belt
segment flexibility kb , and mass mb . The belt friction model we have derived for moving belt and fixed
drum applies equally well to problems involving fixed belt and rotating drum, like band brakes, and to
problems involving both belt and drum rotating, like belt drives. It can be also used to systems with the
belt replaced by a rope or a string.
In case of V-shaped belts (4.16) should be replaced by

Ff = Ft 1 + eµγ·sgn(Ft )/ sin(δ/2)
 
(4.17)
where δ is the angle of the ‘V’ of the belt cross section.
Fig. 4.9a shows an example of a hoisting machine lifting a load by means of a cable passing over an
undriven sheave and wrapped around a driven drum. The rotary motion of the drum drive is transformed
into the translational motion of the cable by means of an ideal transformer with the ratio 1 : r1 , where r1
is the drum radius. The c1 and c2 parts of the cable are modelled by elements characterizing the cable
flexibility and damping. The element parameters of the cable part c2 depend on the cable length ℓ of
c2 . The friction between the cable and the sheave is represented by the belt-friction model given above.
The sheave dynamics, converted from rotary to translational motion along the cable, is represented by
the elements of reduced parameters
mr = Js /r2 , br = bs /r22 (4.18)
where Js is the moment of inertia of the sheave, r2 is its radius and bs is the damping of the sheave
bearing.
4-8 MODULE 4. COUPLED MECHANICAL SYSTEMS

B y
c1
g
C A 0
w
G
F c2

E -l (a)

b1 b2(l) l0
l
F E
DRIVE G k1 B BELT A k (l)
w 1:r1 FRICTION
2 mg
& DRUM
m

C
br
mr (b)

Figure 4.9: Hoisting machine.

4.3 Translatory-to-rotary couplings


4.3.1 Translatory-rotational transformer
Translatory-to-rotary coupling is a coupling between the translational motion and a rotational motion
about a fixed axis. Such couplings can be easily modeled by the pure translatory-rotational trans-
former. It is an idealized mechanical device which determines ratio of the velocities as well as the ratio
of the force and torque associated with the two coupled motions. Fig. 4.10 shows our symbol for pure
transformers. The variables associated with this symbol in Fig. 4.10 correspond to the special case when
the symbol is applied to a rotational transformer.

Figure 4.10: Pure translatory-rotational transformer.

Rotational motion of point A with respect to point A along the x axis is represented by poles Ax
and A x , respectively. Rotational motion of point B with respect to point B rotating about the same
axis with the velocity ϕ̇ is represented by poles Aϕ and A ϕ . F is the force exerted by the transformer
to move points A and A apart while counterbalancing external forces applied to these points. τ is the
torque counterbalancing external torques applied to points B and B .
Relation between angular velocities associated with the rotational transformer is

ẋ/ϕ̇ = n (4.19)

where ẋ is the relative velocity of point A with respect to point A , and ϕ̇ is the relative angular velocity
of point B with respect to point B . n is the transformer ratio. An ideal transformer has this ratio
constant.
4.3. TRANSLATORY-TO-ROTARY COUPLINGS 4-9

The force and torque associated with the translatory-rotary transformer are related as

F/τ = −n (4.20)

The power put into the rotational transformer is

P = F ẋ + τ ϕ̇ (4.21)

After submitting (4.19) and (4.20), (4.21) becomes P = 0. A pure translatory-rotary transformer
is thus an idealized mechanical device which transforms a translational motion into a rotational motion
and vice versa. A pure transformer performs this function with no energy loss or storage, so that the net
power flow into a pure transformer is zero.

4.3.2 Rack-and-pinion
In motion control problems it is often necessary to convert rotational motion into a translational one and
vice versa. For instance, a load may be controlled to move along a straight line through a rotary motor
and rack-and-pinion assembly, such as that shown in Fig. 4.11a. The pinion is a small gear, and the rack
is a linear member with the gear on one side. The describing relation of this system is defined as

x = rϕ (4.22)

tA tA
jA jA tA FB

jA xB
FB FB

(a) (b) (c)

Figure 4.11: (a) Rack-and-pinion gear train, (b) equivalent system, (c) pure model.

Thus the constitutive relation of pure rotary-linear motion converters is


     
ϕ̇(t) 0 n τ (t)
= ·
F (t) −n 0 ẋ(t)

where the coupling ratio


ϕ̇ F 1
n= =− =
ẋ τ r

tA jA tA FA
FA

jA FB xA
FB
xB
xA xB
(a) (b)

Figure 4.12: Movable rack-and-pinion assembly.

To apply this result to the rack-and-pinion assembly movable along the x-axis as that shown in
Fig. 4.12a, the variable x should be replaced by xAB .
4-10 MODULE 4. COUPLED MECHANICAL SYSTEMS

4.3.3 Pulley or sprocket assembly


Another approach to rotary-translatory motion conversion is to utilize the rotary motion of a pulley
combined with the linear motion of a belt. The same effect is achieved by combining a sprocket with a
chain, a drum with a rope or cable wound around it, etc. This concept is outlined in Fig. 4.13a.
A common assembly of two pulleys and a belt used for the control of a mass through one of the pulleys
by a rotary prime mover is shown in Fig. 4.13b. Both the drive and idle pulley have the same radius r.

FB
FB tA FB
tA jA
tA
jA jA xB

(a) (b) (c)

Figure 4.13: (a) Pulley or sprocket. (b) Belt and pulley rotary-translatory assembly.

4.3.4 Lead screws


Fig. 4.14 shows a similar situation in which a lead screw is used as the rotary-to-translatory mechanical
linkage. The lead screw is used there to drive a payload along the z-axis. The screw is fixed free to rotate
and as it is turned, the nut moves along the screw with the payload attached. In this case, the describing
relation is defined as the screw rotation ϕ per unit of linear motion x. This is also referred to as the pitch
P of the screw (the reciprocal of the pitch is called the lead). Therefore, the describing relation is

ϕ = ±P x

and the coupling ratio of the lead screw is simply

n = ±P

The sign is related to the mutual orientation of the rotational and translational motions.

tA FB
jA

jA xB
tA
(a) (b)

Figure 4.14: Lead screw assembly.

4.3.5 Slider cranks


Also various mechanical linkages are used as rotary-translatory converters. As an example of this ap-
proach, Fig. 4.15 shows a configuration of a crank driving a linear stage. The crank of length r rotates
about its center A and has a rod of fixed length l mounted to its rotating end. The other end of the
connecting rod B is attached to a linear stage which is constrained to move in only along the x-axis. As
the describing relation is p
xBA = r cos ϕA + l2 − (r sin ϕA + y0 )2
the reciprocal of the coupling ratio is

1 ẋBA r cos ϕA (r sin ϕA + y0 )


= = −r sin ϕA − p
n ϕ˙A l2 − (r sin ϕA + y0 )2
4.4. LOAD CONVERSIONS 4-11

tA FB
tA
jA xB
FB

(a) (b)

Figure 4.15: Slider crank.

As the crank rotates by 360◦, the linear stage moves in both directions along x. There are two
distinct angular positions of the crank that correspond to full extension and retraction of the stage called
dead-center positions. If this module is used to convert linear to rotary motion the direction of rotation
may go in either way in these positions (this problem is usually overcome by combining the crank with
a flywheel).

4.4 Load conversions

C
E E

F F
D (a ) (b )
Figure 4.16: (a) Cascade interconnection of energy-transfer transformers, (b) one-transformer equivalent.

Fig. 4.16a shows a ‘cascade’ interconnection of energy-transfer transformers of coupling ratios n1 and
n2 . A simple analysis of the total transformation of across and through variables in the interconnection
indicates that the cascade can be replaced by only one transformer with the coupling ratio n1 n2 shown
in Fig. 4.16b.

Figure 4.17: Load ‘seen through’ an energy-transfer transformer.

Fig. 4.17 indicates the way in which series or parallel combinations of physical elements connected
to a pole-pair of an energy-transfer transformer are ‘seen through’ the transformer, or referred to at the
other transformer pole-pair. Z resp. Y is an impedance resp. admittance of the generic element, E resp.
J is an independent source of an across resp. through variable. This is the principle behind utilizing
various transducers and couplings to convert the amount of an actuator load at the price of changed
related across and through variables.
4-12 MODULE 4. COUPLED MECHANICAL SYSTEMS
Module 5

Planar mechanical systems

5.1 Planar motion of particles


5.1.1 Planar curvilinear motion
Recollect that a particle is a body whose physical dimensions are so small compared with the radius of
curvature of its path that the motion of the particle can be approximated as that of a point. A particle
is in a planar motion if its path lies in a single plane. Consider the example of a planar curvilinear path
of a particle shown in Fig. 5.1a where point A represents the particle position at time t.

Figure 5.1: (a) Particle position, (b) velocity, and (c) acceleration vectors.

The particle position can be completely specified by the position vector r measured from a fixed origin
O. At time t + ∆t, the particle is at point A’, the position of which is specified by the vector r + ∆r.
The particle displacement during time ∆t is indicated by the vector ∆r representing the vector change
of position. If an origin O’ is chosen at some location different from O, the position vector r changes, but
∆r remains unchanged. The distance traveled by the particle as it moves along the path from A to A’ is
the scalar length ∆s measured along the curved path. Thus, the vector displacement ∆r is distinguished
from the scalar distance ∆s.
The instantaneous velocity v of the particle is defined as the limiting value of the particle average
velocity between the points A and A’ as the points become closer together and the time interval ∆t
approaches zero. Thus,
∆r
v = lim
∆t→0 ∆t

The direction of ∆r approaches that of the tangent to the path as ∆t approaches zero and, hence, the
velocity v is always a vector tangent to the path as shown in Fig. 5.1b.
The definition of the derivative of a scalar quantity can be extended to include a vector quantity so
that
dr
v= = ṙ
dt
Thus the derivative of a vector is itself a vector having both a magnitude and a direction. The magnitude
of v is called the speed and is the scalar
ds
v = |ṙ| = = ṡ
dt

5-1
5-2 MODULE 5. PLANAR MECHANICAL SYSTEMS

In Fig. 5.1b, the velocity of the particle at A is denoted by the tangent vector v and the velocity at A’
by the tangent vector v′ . There is thus a vector change in the velocity during the time ∆t. The velocity
v at A plus the change ∆v must equal the velocity at A’. So we may write v′ − v = ∆v. Inspection of
the vector diagram in Fig. 5.1c shows that ∆v depends both on the change in magnitude of v and on the
change in direction of v.
The instantaneous acceleration a of the particle is defined as the limiting value of the average accel-
eration between A and A’ as ∆t approaches zero, or
∆v
a = lim
∆t→0 ∆t
By definition of the derivative we write
dv
a= = v̇ = r̈
dt
As the interval ∆t approaches zero, the direction of the change ∆v approaches that of the differential
change dv and, hence, a. It is apparent from Fig. 5.1c that the acceleration a reflects change in magnitude
of v as well as the change of direction of v. In general, the direction of the acceleration of a particle in
curvilinear motion is neither tangent nor normal to the particle path.

5.1.2 Planar motion in rectangular coordinates


Absolute planar motion
Fig. 5.2a shows the curvilinear path of a particle represented by point A along with the rectangular
coordinate axes x and y. Such a coordinated system, drawn on the plane of the particle motion, is often
called Cartesian. The particle motion can be described at any time t by specifying its position xA and
yA measured from the origin of the x- and y-axes O. If the coordinate system is fixed to the absolute
frame the rectangular components indicate the absolute motion of the particle.

Figure 5.2: Rectangular components of (a) of position, velocity, and (b) acceleration.

The position vector rA , the velocity vA , and the acceleration aA of the particle are represented in
Fig. 5.2 together with their x- and y-components. Using the unit vectors i and j, we may write the vectors
rA , vA , and aA as
rA = xA i + yA j
vA = ṙA = vAx i + vAy j (5.1)
aA = v̇A = r̈A = aAx i + aAy j
The unit vectors have no time derivatives since their magnitudes and directions remain constant. Thus
the scalar values of the components of vA and aA are just vAx = ẋA , vAy = ẏA and aAx = v̇Ax = ẍA ,
aAy = v̇Ay = ÿA .
It is obvious from Fig. 5.2 that the absolute values vA = |vA | and a = |aA | are related to the x- and
y-components as q q
vA = 2 + v2
vAx aA = a2Ax + a2Ay
Ay

Relations (5.1) prove that the planar motion of a particle can be decomposed into motions in two
different directions (mutually perpendicular in this case). Practical experience indicates that the opposite
procedure – vector composition of planar motions – is also possible. Such principle of motion composition
is considered as an axiom, it cannot be proven.
Thus, if the coordinates x and y of a particle are known as functions of time, x = fx (t) and y = fy (t),
then for any value of the time we may combine them to obtain r. Similarly, we combine their first
5.1. PLANAR MOTION OF PARTICLES 5-3

derivatives ẋ and ẏ to obtain v and their second derivatives ẍ and ÿ to obtain a. Or, if the acceleration
components ax and ay are given as functions of the time, we may integrate each one separately with
respect to time, once to obtain vx vy and again to obtain x = fx (t) and y = fy (t). Elimination of the
time t between these last two parametric equations gives the equation of the curved path y = f (x).

Relative planar motion


Consider now two particles positioned at points A and B having separate curvilinear motions in a given
plane with the x-y axes, Fig. 5.3a. The absolute positions of the particles are determined by the vectors
rA and rB measured from the origin O. The relative position of A measured with respect to the position
of B is represented by the vector
rAB = rA − rB (5.2)

Figure 5.3: Relative motion (a) of A with respect to B, and (b) of B with respect to A.

Using the unit vectors i and j along the x- and y-axes we can write

rAB = xAB i + yAB j (5.3)

where xAB = xA − xB and yAB = yA − yB are the x- and y-components of the relative position of A with
respect to B.
If we now differentiate (5.2) and (5.3) with respect to time we obtain the relative velocity of A as
observed from B
vAB = ṙAB = ṙA − ṙB
where vABx = vAx − vBx and vABy = vAy − vBy are the x- and y-components of the relative velocity of A
with respect to that of B. These can be expressed in terms of the rectangular components of the absolute
velocities vAx , vBx , vAx , vBy of points A and B.
If we differentiate (5.2) and (5.3) with respect to time twice we will get the relative acceleration of A
with respect to B as
aAB = ȧAB = ṙA − ȧB = aABx i + aABy j
where the x- and y-components aABx = aAx − aBx and aABy = aAy − aBy .
Let us now exchange the role of the two points to observe the motion of B as seen from A. Fig. 5.3b
indicates that the corresponding expressions for the relative position, velocity, and acceleration become

rBA = rB − rA = −rAB
vBA = vB − vA = −vAB
aBA = aB − aA = −aAB

Postulate of compatibility of planar motions


Due to geometric continuity of space the postulate of compatibility of motions holds also for planar
motions. Let us consider curvilinear motions of points A, B, and C shown in Fig. 5.4. The vectors rAB ,
rBC , and rCA indicate the mutual relative motions of the points.
Note that if the vectors are oriented in the same direction along the loop that they form

rAB + rBC + rCA = 0


5-4 MODULE 5. PLANAR MECHANICAL SYSTEMS

Figure 5.4: Composition of relative positions in a plane.

or
(xAB + xBC + xCA )i + (yAB + yBC + yCA )j = 0
when the vectors are decomposed into their x- and y-components. Thus postulate of motion compatibility
applies independently also to the x- and y-components

xAB + xBC + xCA = 0 and yAB + yBC + yCA = 0

The postulate of motion compatibility holds for velocities and accelerations of the three points:

vAB + vBC + vCA = 0


aAB + aBC + aCA = 0

In general, the postulate of motion compatibility holds for any motions along a closed loop passing
through the sequence of an arbitrary number of points. This trivial-looking statement plays an important
role in the formulation of equations describing the behavior of mechanical systems.

5.1.3 Planar forces in rectangular coordinates


Composition of planar forces
A force has been defined as the action of one body on another. We find that force is a vector quantity,
since its effect depends on the direction as well as on the magnitude of the action and since forces may
be combined according to the parallelogram law of vector combination. Two forces F1 and F2 that are
concurrent may be added by the parallelogram law in their common plane to obtain their sum or resultant
R as shown in Fig. 5.5.

Figure 5.5: Composition of planar forces.

In addition to the need for combining forces to obtain their resultant, we often have occasion to
replace a force by its vector components which act in specified directions. By definition, the two or more
vector components of a given vector must vectorially add to yield the given vector. Thus, the force R
in Fig. 5.5 may be replaced by or resolved into two vector components F1 and F2 with the specified
directions merely by completing the parallelogram as shown to obtain the magnitudes of F1 and F2 .
Let us now consider a particle subjected to the action of concurrent forces F1 , F2 , F3 , ... assuming
that the lines of action of all the forces are situated in the same plane as the trajectory of the particle.
The vector sum of the applied forces becomes
X
Fa = Fi (5.4)
i

The sum must include all the external forces applied to the particle in question. The reliable way to
account accurately and consistently for every force is to form the free-body diagram of the particle under
5.1. PLANAR MOTION OF PARTICLES 5-5

consideration isolating it from all contacting and influencing bodies and replacing the bodies removed by
the forces they exert on the particle isolated.
Fig. 5.6a gives an example of three forces applied to a particle represented by point A while Fig. 5.6b
shows the resultant Fa of the applied forces.

Figure 5.6: (a) Forces applied to a particle, (b) vector and (c) orthogonal components of the resultant
external force.

The vector of the total applied force Fa is shown again in Fig. 5.6c where it is decomposed along the
x and y coordinate axes. With the aid of the unit vectors i and j, we may rewrite (5.4) in terms of the
x- and y-components of the individual applied forces as

Fa = Fax i + Fay j

where X X
Fax = Fxi and Fay = Fyi
i i

are the x- and y-components of the resultant force applied to the particle.
The absolute value Fa = |Fa | is
sX X
Fa = ( Fxi )2 + ( Fyi )2
i i

Equilibrium of planar forces


If a particle of mass m is isolated in the inertial system and is subjected to the action of a sum of applied
forces Fa then Newton’s second law of motion may be described by the vector relation

Fa = ma (5.5)

where the acceleration a of the particle is always in the direction of the applied force Fa . This is shown
in Fig. 5.7a where the particle is represented by point A. This effect is the inertia of the particle mass
which is its resistance to rate of change of velocity. Equation (5.5) is usually referred to as the equation
of motion in the vector form.

Figure 5.7: (a) Particle acceleration due to applied force, (b) force dynamic equilibrium.

Fig. 5.7b suggests that the inertial effect of the particle mass can be replaced by the hypothetical
inertial force Fm = −ma acting on the massless point representing the particle. Then (5.5) can be
written in the alternative form
Fa + Fm = 0 (5.6)
interpreted as the dynamic equilibrium of forces acting at on the particle.
5-6 MODULE 5. PLANAR MECHANICAL SYSTEMS

Relation (5.6), known as d’Alembert’s principle, has far-reaching implications in dynamic analysis. It
allows us to treat the particle inertial force Fm in the same way as the external forces applied to the
particle.
If a particle is subjected to the action of a sum of applied forces Fa and its velocity with respect to
the inertial system is zero or constant
Fa = 0
This expresses the static equilibrium of forces acting at on the particle.

5.1.4 Dynamic model of particles in planar motion


From the foregoing discussion it should be recognized that the rectangular-coordinate representation of
curvilinear motion is merely the superposition of the components of two simultaneous rectilinear motions
in the x- and y-directions. Therefore, everything covered so far on rectilinear motion may be applied to
the x-motion and to the y-motion.

Figure 5.8: (a) Submodel for particle motion in the vertical plane, (b) submodel symbol.

Fig. 5.8b gives the three-pole dynamic model of a particle pursuing a curvilinear motion in the vertical
x-y plane which is shown in Fig. 5.8a. The particle position in the x-y plane denotes point Ax. Poles
Ax and Ay represent point velocities in the horizontal and vertical directions. The inertia effects of
the particle mass m in the x- and y-directions are modelled by inertors. A source of force represents
the gravitational force Fg = g m acting on the particle in the vertical direction, g is the gravitational
acceleration.
The third pole corresponds to the absolute frame representing the velocity reference. Fig. 5.8c shows
the symbol of the submodel with pins corresponding to the poles Ax and Ay. The third pole associated
with the velocity reference and respected in the submodel does not need to be shown explicitly in the
symbol.
Obviously, also a multipole model of a particle in a curvilinear motion can be decomposed into two
rectilinear models, one for dynamic interactions in the x direction, and the other one in the y direction.
In most practical cases these two models will be coupled, however.

Example: Cannon ball


Let us investigate the trajectory of a cannon ball launched from the coordinate system origin at t = 0
with the initial speed v(0) at the angle α = 45◦ . Fig. 5.9a shows the expected parabolic trajectory in the
vertical plane if the cannon ball represented by point B.

Figure 5.9: (a) Trajectory of a cannon ball, (b) ball dynamic model.

To model the ball dynamics, we apply the submodel shown in Fig. 5.8 augmented by blocks for
computing the ball xB and xB position by integrating the ball velocities
√ Bx and By . The√ x- and y-
components of the ball initial speed v(0) are obviously vx (0) = v(0)/ 2 and vy (0) = v(0)/ 2.
5.2. PLANAR SYSTEMS WITH MASSLESS RODS 5-7

We can see that the ball x- and y-motions are mutually independent under the simplifying consider-
ation. If we had introduced, for example, aerodynamic drag force exerted on the ball which depends on
the speed squared, then the x- and y-parts of the ball model would be coupled. As a result, the ball x-
and y-motions would be interdependent, and the trajectory would not be parabolic any more.

5.2 Planar systems with massless rods


5.2.1 Rod with revolute joints
Model in rectangular coordinates
Motion of particles A and B in the x-y plane is constrained by a rigid rod of negligible mass as shown in
Fig. 5.10a. The rod configuration can be characterized by the relation

x2AB + yAB
2
= ℓ2 (5.7)

where ℓ is the rod length whereas xAB = xA − xB and yAB = yA − yB are projections of the rod onto the
x and y axes. After differentiating (5.7) with respect to time we get the geometric constraint imposed by
the rod on the particle motion
xAB ẋAB + yAB ẏAB = 0 (5.8)

Figure 5.10: (a) Particles interconnected by massless rod, (b) transformer model.

The rod must be in equilibrium, i.e. all forces and moments applied to it must be in balance. The
free-body diagram of the rod in Fig. 5.10a indicates that the balance of forces can be written in the form

FAx + FBx = 0 FAy + FBy = 0 (5.9)

end the balance of moments acting on the rod can be expressed as

(FAy − FBy ) xAB − (FAx − FBx ) yAB = 0

After substituting for FAx and FAy from (5.9), the previous relation becomes

FAy xAB − FAx yAB = 0 or FBy xAB − FBx yAB = 0 (5.10)

Let us now arrange (5.8) and (5.10) in a common matrix form


     
ẋAB 0 n FBx
= · (5.11)
FBy −n 0 ẏAB

where n = −yAB /xAB .


This result leads us to the conclusion that the constraint provided by the massless rod can be modeled
by the pure transformer with the variable ratio n as shown in Fig. 5.10b.
An alternative but dynamically equivalent model characterized by the symbol shown in Fig. 5.11a is
given in Fig. 5.10b. Values of variables are interrelated in the model by equations (5.9) and (5.10).

Example: Mathematical pendulum


yAB
n=
xAB
5-8 MODULE 5. PLANAR MECHANICAL SYSTEMS

Figure 5.11: (a) Massless rod, (b) its model.

Figure 5.12: (a) Mathematical pendulum, (b) its dynamic model.

Example: Double mathematical pendulum


yB yC − yB
n1 = n2 =
xB xC − xB

Example: Carriage with pendulum


yAB
n=
xAB

Example: Sliding rod


yAB
n=−
xAB

Model in polar coordinates


There are system models in which we need to observe or control the angle or torque of a rod. If this
is the case we can apply to the rod one of the models shown in Fig. 5.16 or Fig. 5.17. The models are
related to the rod geometric configuration shown in Fig. 5.16a.
Positions of the rod end-points A and B are constrained as

xAB = ℓ cos ϕ yAB = ℓ sin ϕ (5.12)

where ℓ is the rod length. Differentiation with respect to time gives the velocity constrains

ẋAB = −ℓ sin ϕ · ϕ̇ ẏAB = ℓ cos ϕ · ϕ̇ (5.13)

Balance of forces can be written in the form

FAx + FBx = 0 FAy + FBy = 0 (5.14)

Balance of moments related to the rotation of A about B

M − FAx yAB + FAy xAB = 0 (5.15)

where M is an external moment of force applied to the rod.


After substituting for xAB and yAB from (5.12) and setting M = Mx + My where

Mx = −FBx ℓ sin ϕ My = FBy ℓ cos ϕ


5.2. PLANAR SYSTEMS WITH MASSLESS RODS 5-9

B n1
C xB
x m1 By
0 xB xC n2 yB
y C2

m1 m 1g
Cx xC
Cy
yC
m2
m2 m 2g
(a) (b)

Figure 5.13: (a) Mathematical double pendulum, (b) its dynamic model.

Figure 5.14: (a) Carriage with pendulum, (b) its dynamic model.

Combining now these relations with (5.13) we can write


           
ẋAB 0 nx FBx ẏAB 0 ny FBy
= · = · (5.16)
Mx −nx 0 ϕ̇ My −ny 0 ϕ̇
where
nx = −ℓ sin ϕ ny = ℓ cos ϕ
Obviously, the resulting model of the rod can be composed from two pure transformers as shown in
Fig. 5.16b.
An alternative model combining symbols and equations is in Fig. 5.17. The sources of force and torque
are interrelated by equations (5.12), (5.13)(5.14), and (5.28).

Example: Slide-crank mechanism

5.2.2 Rod with revolute and translational joints


Model in rectangular coordinates
Motion of particles A and B in the x-y plane is constrained by a rigid rod of negligible mass as shown in
Fig. 5.19a. The rod configuration can be characterized by the relation
x2AB + yAB
2
= ρ2AB (5.17)
where ρAB is distance of point A measured from point B along the rod. Differentiation of (5.17) with
respect to time gives the geometric constraint
xAB ẋAB + yAB ẏAB = ρAB ρ̇AB (5.18)
5-10 MODULE 5. PLANAR MECHANICAL SYSTEMS

Figure 5.15: (a) Sliding rod, (b) its dynamic model.

Figure 5.16: Massless rod in polar coordinates.

Let us rewrite this as


ρ̇AB = ρ̇x + ρ̇y (5.19)
where
xAB yAB
ρ̇x = ẋAB ρ̇y = ẏAB (5.20)
ρAB ρAB
The free-body diagram of the rod in Fig. 5.19a indicates that the balance of forces can be written in
the form
FAx + FBx = 0 FAy + FBy = 0 (5.21)
and the balance of moments as

FAx ρAB + Fρ xAB = 0 FAy ρAB + Fρ yAB = 0 (5.22)

Combining now these relations with (5.18) we can write


           
ρ̇x 0 nx Fρ ρ̇y 0 ny Fρ
= · = · (5.23)
Fx −nx 0 ẋAB Fy −ny 0 ẏAB

where
nx = xAB /ρAB ny = yAB /ρAB
Obviously, the resulting model can be composed from two pure transformers as shown in Fig. 5.19b.

Model in polar coordinates


As shown in Fig. 5.20a, positions of points A and B are constrained as

xAB = ρ cos ϕ yAB = ρ sin ϕ (5.24)

where ρ is the distance of A from B. Differentiation with respect to time gives the velocity constrains

ẋAB = ρ̇ cos ϕ − ρϕ̇ sin ϕ ẏAB = ρ̇ sin ϕ + ρϕ̇ cos ϕ (5.25)

Balance of forces can be written in the form

FAx + FBx = 0 FAy + FBy = 0 (5.26)


5.2. PLANAR SYSTEMS WITH MASSLESS RODS 5-11

Figure 5.17: Massless rod in polar coordinates: (a) graphical symbol, (b) model diagram.

Figure 5.18: (a) Slide crank mechanism, (b) model diagram.

and
FAx cos ϕ + FAy sin ϕ = 0 (5.27)
Balance of moments is
τ − FAx yAB + FAy xAB = 0 (5.28)
where τ is an external moment of force applied to the rod.
In this case, the resulting model can be composed from four pure transformers as shown in Fig. 5.20b
where
nxϕ = −ρ sin ϕ nyϕ = ρ cos ϕ nxρ = cos ϕ nyρ = sin ϕ

5.2.3 Transformations
An example of the magnitude of the error introduced by neglect of the motion of the earth may be cited
for the case of a particle that is allowed to fall from rest (relative to the earth) at a height h above
the ground. We can show that the rotation of the earth gives rise to an eastward acceleration (Coriolis
acceleration) relative to the earth and, neglecting air resistance, that the particle falls to the ground a
distance s
2 2h3
x= ω cos γ
3 g
east of the point on the ground directly under that from which it was dropped. The angular velocity of
the earth is ω = 0.729(10−4) rad/s, and the latitude, north or south, is γ. At a latitude of 45◦ and from
a height of 200 m, this eastward deflection would be x = 43.9 mm.
These corrections are negligible for most engineering problems which involve the motions of structures
and machines on the surface of the earth. In such cases, the accelerations measured with respect to
reference axes attached to the surface of the earth may be treated as ”absolute,” and Eq. 3/2 may be
applied with negligible error to experimental measurements made on the surface of the earth.* If the ideal
experiment described were performed on the surface of the earth and all measurements were made relative
to a reference system attached to the earth, the measured results would show a slight discrepancy upon
substitution into Eq. 3/2. This discrepancy would be due to the fact that the measured acceleration
would not be the correct absolute acceleration. The disc The discrepancy would disappear when we
introduced the corrections due to the acceleration components of the earth.

5.2.4 Planar motion in polar coordinates

x = r cos ϕ y = r sin ϕ (5.29)


where r is the length of vector r and ϕ is the angle between this vector and the axis x.
By differentiating these relation with respect to time we will obtain velocities of point A in the
coordinate system x, y expressed in terms of velocities in the coordinate system ρ, ϕ
5-12 MODULE 5. PLANAR MECHANICAL SYSTEMS

Figure 5.19: (a) Rod in rectangular coordinates, (b) model.

Figure 5.20: (a) Rod in polar coordinates, (b) model.

vx = vρ cos ϕ − vϕ sin ϕ (5.30)


vy = vρ sin ϕ + vϕ cos ϕ (5.31)

Fρ = −Fx cos ϕ − Fy sin ϕ (5.32)


Fϕ = +Fx sin ϕ − Fy cos ϕ (5.33)

     
vxy 0 N Fxy
= · (5.34)
Fρϕ −Nt 0 vρϕ
where
       
vx vρ Fx Fρ
vxy = vρϕ = Fxy = Fρϕ = (5.35)
vy vϕ Fy Fϕ
and
 
cos ϕ − sin ϕ
N= (5.36)
sin ϕ cos ϕ

5.2.5 Planar motion of particles in path coordinates


The motion of A may also be described by measurements along the tangent t and normal n to the curve.
The direction of n lies in the local plane of the curve. These last measurements are known as path
variables.
Module 6

Electrical systems

6.1 Electrical interactions


6.1.1 Electrical voltage and current
The electrical voltage [V] (or potential difference) vBA between two points such as B and A is defined
as the work which would be done (or the energy required) in carrying a unit positive charge from one
point to the other. For the concept of voltage to be applicable, it is necessary that the work determined
in this way be independent of the path along which the charge is moved. This requirement implies that
the rate of change of the magnetic field in the region of the path must be either zero or negligible.
This requirement fits into the framework of the quasi-static field theory which implies that the rate
of change of magnetic field normal to the region of the path on the subsystem energy boundary must be
negligible.
If it is necessary to do work on a positive charge (i.e., to apply a force in the direction of motion) as
it moves from point A to point B, then point B is said to be at a higher potential than point A and the
voltage vBA is considered positive.
Since it is often not known a priori which of two points will be at the higher potential or voltage, it is
necessary to establish an algebraic reference convention for voltage. In this course, the assumed positive
orientation of voltage is denoted by an empty-head arrow pointing in the direction of the from the point
of a lower to the point of a higher potential.
Electrical current i [A] is defined as the rate of flow of charge across a given area, often the cross-
sectional area of a wire. As the current may be changing with time, it is
dq
i(t) = (6.1)
dt
where q = q+ − q− is the net charge of the particles crossing the entry, q+ and q− are charges of the
particles charged positively and negatively, respectively. The physical unit of electrical charge – coulomb
[C] – is equal in magnitude but opposite in sign to the charge of 6.22 · 1018 electrons.
Current has a direction as well as a magnitude. The direction convention commonly chosen is that
direction in which a net flow of positive charge has occurred. To establish in a reference for the direction
of a current we will use full-head arrows indicating the direction of the assumed positive current.

6.1.2 Electrical power, energy and work


From the definition of voltage, the amount of work dWBA done when an infinitesimal charge dq is moved
from A to B is dWBA = vBA dq, or
dWBA
vBA = (6.2)
dq
One joule [J] is equivalent to one watt-second [W.s].
The electrical power P delivered to a two-terminal electrical element can be computed from the
definitions of voltage and current, (6.2) and (6.1). For the element

dWBA dq dWBA
P= = = i vBA (6.3)
dt dt dq

6-1
6-2 MODULE 6. ELECTRICAL SYSTEMS

where i is the current flowing through the element and vBA is the voltage difference across it. In the
MKS system of units, power is measured in watts. One watt is equal to one joule/sec or one N.m/sec.
The electrical energy E supplied to an element is the time integral of the power
Z tb
E= i vBA dτ (6.4)
ta

where ta and tb are the beginning and end of the time interval during which power flows. If energy is
delivered to an electrical system through more than one pair of terminals, the total energy supplied is
the sum of the energies supplied at all the terminals.
The law of conservation of energy, or the first law of thermodynamics, states that the energy (of
all forms) which is delivered to a system must either be stored in the system or transferred out of the
system.

6.2 Prescribed electrical interactions


6.2.1 Prescribed electrical voltages
If an electric circuit is driven by a prescribed voltage v(t) [m/s], the actuator driving the body can be
represented in the circuit model by a pure source of electrical voltage. This two-pole physical element
is characterized by the constitutive relation
v(t) = f (t) (6.5)
where i(t) is the known single-value function of time.

Figure 6.1: Symbols of pure source of electrical (a) voltage and (b) current.

Fig. 6.1a gives our graphical symbol representing a pure source of electrical voltage v(t) in multipole
diagrams. The two short line segments sticking out of the symbol are called pins. They denote poles of
the source representing terminals of the modelled real source of electrical energy. The symbol polarity
is indicated by the + sign. Besides the prescribed electrical voltage v(t), the pure source of voltage is
also associated with its current i(t). This current is imposed on the source by the system in which the
source operates. The assumed positive orientation of v(t) and i(t) in relation to the symbol polarity is
shown in Fig. 6.1a by the empty-head and full-head arrow, respectively.
The chosen orientation of source variables is such that the power of the source corresponds to the flow
of energy into the source. Sources are usually used to model supplies of energy into systems. Despite of
this, we shall always use the same ”consumer” variable-orientation convention for all the physical elements
regardless if they consume or deliver energy. Such a unified approach is for the benefit of modeling and
analysis simplicity. Note, that if the source symbol is oriented properly in a multipole diagram there is
no need to denote in it the variable orientation by arrows.

6.2.2 Prescribed electrical currents


To model energy sources of prescribed electrical currents we need to introduce pure sources of electrical
current. In general, such a source exhibits current i(t) characterized by the constitutive relation
i = f (t) (6.6)
where i(t) is a known single-value function of time.
Our graphical symbol for the source of electrical current is given in Fig. 6.1b. The full-head and
empty-head arrows indicate the assumed positive orientation of the electrical current i(t) and electrical
voltage v(t) associated with the source. The variable orientation is related to polarity of the source
symbol denoted by the + sign. The relative electrical voltage v(t) of the source represents the electrical
voltage of the real electrical energy source determined by the system comprising it.
6.3. PURE ELECTRICAL RESISTORS AND CONDUCTORS 6-3

6.2.3 Indicators of electrical voltages and electrical currents


Pure sources of zero electrical current can be used as physical elements modeling ideal indicators of
electrical voltage, i.e. instruments measuring relative electrical voltage between distinct sites of inter-
action. Our graphical symbol for such an indicator is given in Fig. 6.2a. Direction of the arrow showing
the assumed positive polarity of the indicated electrical voltage is fixed to the polarity of the symbol.

Figure 6.2: Symbol of ideal indicator of electrical (a) voltage and (b) current.

Pure sources of zero electrical voltage can be utilized as physical elements modeling ideal measuring
instruments or indicators of electrical current. Our graphical symbol for such an indicator is given in
Fig. 6.2b. Direction of the arrow showing the assumed positive polarity of the indicated electrical current
is fixed to the polarity of the symbol.
Note the difference between measuring electrical voltage and electrical current. The direct measure-
ment of electrical currents requires disconnecting the system and connecting it back by a electrical current
indicator. Electrical voltages are measured by indicators applied to systems without any need to dis-
connect them. This is the reason why electrical voltages represent electrical across variables while
electrical currents are electrical through variables.

6.3 Pure electrical resistors and conductors


All ordinary materials exhibit some resistance to the flow of electric charge. Resistors are encountered
in many shapes and forms, often designed to provide a needed resistance effect in an electric system, but
sometimes as an unwanted (parasitic) effect. Fig. 6.3a shows a real electrical component called resistor
formed as a coil of resistive wire wound around a ceramic material.
A pure electrical resistor is defined by a single-valued functional relationship between voltage difference
across the resistor v and its current i, such that v = 0 when i = 0, and the signs of v and i are the same.
An ideal electrical resistor is characterized by the linear relation, known as the Ohm’s law,

v(t) = Re · i(t) (6.7)

where the constituency parameter Re [Ω] is called electrical resistance.


In some models it is useful to use an pure or ideal electrical conductor instead of a pure resp. ideal
resistor. Ideal conductors are characterized by electrical conductance Ge reciprocal to Re , i.e. Ge = 1/Re
[S].

Figure 6.3: (a) Real electrical resistor. Pure electrical (b) resistor and (c) conductor. (d) Model of real
resistor.

Fig. 6.3b gives our symbol for pure electrical resistors while in Fig. 6.3c is our symbol for pure
conductors. Polarity of both symbols is denoted by the + sign, and the assumed positive orientation
of the element voltage v and current i are associated with the symbol polarities. In Fig. 6.3d, the pure
conductor is used to model the real resistor shown in Fig. 6.3a. Note that we do not need to show in
the model the orientation of variables by arrows thanks to the unambiguous relation between the symbol
polarity and the variables.
Electric and magnetic field effects are assumed to be zero in pure resistors and conductors. These
elements can store neither electric-field nor magnetic-field energy. Characteristics of real components
6-4 MODULE 6. ELECTRICAL SYSTEMS

called conductors or resistors utilized in electrical and electronic systems, however, comply often quite
well with (6.7) in a fairly wide range of voltages and currents.
For example, if a voltage changing slowly enough is applied across a straight piece of a wire, the
electrical current passing through the wire can be approximated by (6.7), and the electrical conductance
of the wire
A
Ge = γ (6.8)

where γ [S.m−1 ] is the specific conductivity of the material from which the wire is made, A is the
cross-section area of the wire, and ℓ is the wire length.
Electrical power is delivered to a resistance whenever current flows through it. For an ideal resistance
P = Re i2 = v 2 /Ge (6.9)
Power always flows into the resistor regardless of the signs of i and v. The power transferred to a
resistance cannot be retrieved and is said to be dissipated. Usually the power is transformed into thermal
internal energy in the resistance and the temperature rise which accompanies the increase in internal
energy results in flow of heat to the environment.

6.4 Electrical capacitance


In a dielectric material an electric field can be established without allowing a significant flow of charge
through it. If two pieces of conducting material are separated by a dielectric material, an electric field
is established between the conductors when charge flows into one conductor and out of the other. This
electric field results in a potential difference between the two conductors which depends on the amount
of charge placed on the conductors. Physical devices which exhibit this type of relation between charge
and voltage are said to have capacitance.

Figure 6.4: (a) Parallel-plate capacitor. (b) Pure electrical capacitor. (c) Model of real capacitor.

Most electrical components exhibit the property of capacitance to some extent. A device designed
to behave primarily like a capacitance is called an electrical condenser or capacitor. As an example,
Fig. 6.4a shows a real capacitor consisting of two parallel flat plates separated by a dielectric.
To permit this capacitive effect to be separated from other electrical phenomena, we shall define a
pure capacitor. It is a two-terminal electrical element in which charge q is a function of the voltage
difference across the element:
q = f (v)
In a pure capacitor, the conductors offer zero resistance to flow of charge, the dielectric separating the
conductors has infinite resistance to the flow of charge, and magnetic field effects are absent.
Fig. 6.4b shows the symbol to be used for a pure capacitor. The assumed positive direction of current
flow i and voltage drop v is indicated by an arrows drawn beside the symbol. This sign convention is
related to the polarity of the symbol indicated by the + sign. In Fig. 6.4c, the pure capacitor is used to
model the real capacitor. Note the relation between the orientation of the symbol and the variables in
the model.
The constitutive relationship for an ideal capacitor has charge proportional to voltage difference
q = C ·v
where C is the capacitance of the element, having units of farads [F] (F = A.s/V ). The elemental
equation for an ideal capacitor of constant C is therefore
dv
i(t) = C (6.10)
dt
6.5. PURE ELECTRICAL INDUCTOR 6-5

or
t
1
Z
v(t) = i(τ ) dτ + v(t0 ) (6.11)
C t0

where v(t0 ) is the voltage across the capacitance when t = 0.


For example, it may be recalled from physics that the capacitance of the closely spaced parallel-plate
capacitor shown in Fig. 6.4a is
C = ǫA/d
where C [F] is the capacitance, ǫ [F/m] is the permittivity (a property) of the medium between the
plates in farads/meter, A [m2 ]is the plate area, and d [m] is the plate spacing. For air, for example,
ǫ = 8.85 × 10−12 [F/m].
The capacitance of various real structures depends only on the geometry and the material properties.
In general, the calculation of capacitance for particular structures requires advanced methods of electric-
field theory. Of course the capacitance of an existing element can always be determined experimentally.
When charge is caused to flow into a capacitance, electrical energy Ee is transferred to the element.
From the power equation Pe = i.v, the definition of current i = dq/dt, and the defining equation for a
pure capacitor (6.11), the energy stored in an ideal capacitance is obtained from (2-79) with q = Cv21:
Z v
1 1 q2
Ee = C v, dv = Cv 2 =
0 2 2C
The process of energy storage is reversible, and all the electrical energy stored in an ideal capaci-
tor is retrievable. Note that the energy stored depends neither on the sign of the voltage nor on the
instantaneous value of the current:
Ee ≥ 0.

6.5 Pure electrical inductor


When current flows through a conductor, a magnetic field is established in the space or the material
around the conductor. If this current is changed as a function of time, the intensity of the magnetic field
will also vary with time. According to Lenz’s law, this changing field will induce voltage differences in
the conductor which will tend to oppose the changing current.

Figure 6.5: (a) Electrical coil. (b) Pure electrical inductor. (c) Coil model.

An electrical element is said to be a pure electrical inductor provided that the magnetic flux Φ is
a function of the current i in the element generating the flux,

Φ = Φ(i) (6.12)
The function Φ(i) depends on the magnetic properties of the medium in the vicinity of the current flow.
By winding a wire in a geometry such as the helical coil of Fig. 6.5a, it is possible to reinforce the
flux inside the coil and to have the same flux link several turns of the coil. If a flux Φ links N turns, the
flux linkage λ is
λ = N Φ = f (i) (6.13)
The unit of flux linkage is weber [W] (1 W = 1 V.s).
As shown in Fig. 6.6, the flux linkage λ can be proportional to i, or it can be highly nonlinear
depending if the medium surrounding the inductor is, or is not, nonmagnetic.
The relationship of the magnetic flux linkage to terminal voltage of an inductor is given by Faraday’s
law
dλ(t)
v(t) = (6.14)
dt
6-6 MODULE 6. ELECTRICAL SYSTEMS

Figure 6.6: Constitutive characteristics of pure inductors.

A voltage is ‘induced’ in the inductor which is equal to the rate of change of the flux linkages in the
inductor. Substituting (6.13) into (6.14), the terminal voltage can be written

dλ di
v(t) = · (6.15)
di dt
A linear pure electrical inductor, called an ideal electrical inductor, has flux linkage proportional
to its current:
λ(t) = Le · i(t) (6.16)
The electrical inductance Le of the ideal inductor is measured in henries [H] (1 H = 1 V.s/A), and
λ is defined to be zero when i = 0. As with capacitance, the computation of the inductance for a given
structure will usually require the use of field theory. It is noted that inductance is determined purely by
geometry and material properties.
For example, the inductance of a closely wound, long helical coil as that shown in Fig. 6.5a can be
estimated as
πN 2 µd2
Le =
4l
where l is the coil length, d is its diameter, N is the number of its turns, and µ is the permeability of the
material in the vicinity of the coil. For a vacuum (or very nearly for air) µ = µ0 = 4 × 10−7 H/m.
The inductance of a coil in air will be nearly linear, but the magnitude of the inductance is very
small. This magnitude may be increased by factors up to 106 by the use of ferromagnetic material such
as iron, cobalt, and nickel. However, inductors which use ferromagnetic materials are nonlinear and
give responses somewhat like the nonlinear curve shown in Fig. 6.6. The flattening of this curve as i is
increased is called saturation.
Inductance of a toroidal coil made of a conducting wire wound around a toroid made of a nonmetallic
material such as wood is approximately

N 2A
L e = µ0 [H] (6.17)
l
where N is the number of turns, A is the cross-sectional area, and l is the midcircumference along the
toroid.
The elemental equation for an ideal inductor of constant Le follows from (6.14) and (6.16):

d
v(t) = Le i (6.18)
dt
or
t
1
Z
i(t) = v(τ ) dτ + i(t0 ) (6.19)
Le t0

where i(t0 ) is the inductor current when t = t0 .


Fig. 6.5b gives our symbol of pure electrical inductors. In Fig. 6.5c, the pure inductor is used to model
the helical coil.
6.6. COUPLED INDUCTORS 6-7

An inductor stores energy in the magnetic field associated with the current. The electrical energy
stored in a pure inductor is called the magnetic-field energy
Z tb Z λb
Em = v idτ = idλ (6.20)
ta λa

The energy Em is represented by the shaded area in Fig. 6.6.


If the inductor is ideal, equations (6.16) and (6.20) give
1 1 λ2
Em = Le i2 = (6.21)
2 2 Le
This energy is stored by virtue of the current and does not directly depend on the voltage difference.
The energy is always positive,
Em ≥ 0
and any energy stored by increasing the current from ia to ib is recovered completely when the current
is decreased to ia .

6.6 Coupled inductors


In this unit we shall pay a special attention to electrical inductors mutually coupled by a common
magnetic field. The coupled inductors are of crucial importance as models in many practical applications,
like in electrical transformers, motors and generators. In most cases, such devices are modelled ‘from
the electrical point of view’ without the necessity (or possibility) of modeling detailed structure of the
magnetic field inside the device.
Fig. 6.7a shows a pair of coils placed close to one another so that there is an interaction between their
magnetic fields. The magnetic flux in one coil depends on the currents passing through both coils. Let
us assume that the magnetic flux φ1 of the first coil, and the magnetic flux φ2 of the second coil, can be
expressed as
φ1 (i1 , i2 ) = φ11 (i1 ) + φ12 (i2 )
φ2 (i1 , i2 ) = φ21 (i1 ) + φ22 (i2 ) (6.22)

Figure 6.7: (a) Pair of interacting coils, (b) coupled electrical inductors.

For the following discussions, it will be more convenient to consider magnetic linkages of the coils
instead of their fluxes. Recall, that if all the N1 turns of the first coil are linked by the flux φ1 , the flux
linkage of the first coil is λ1 = N1 φ1 . Similarly, if all the N2 turns of the second coil are linked by the
flux φ2 , than λ2 = N2 φ2 .
Assuming further, that the magnetic fluxes are linear functions of the coil currents, we can express
the magnetic linkages of the two coils as
λ1 (t) = L11 i1 (t) + L12 i2 (t)
λ2 (t) = L21 i1 (t) + L22 i2 (t) (6.23)
where L11 and L22 , is the self-inductance of the first and second inductor, respectively. L12 = L21 are
the mutual inductances of the inductors. Thus the two coils can be modelled by a pair of coupled linear
electrical inductors shown in Fig. 6.7b. Their constitutive relation (6.23) rewritten in matrix form is
     
λ1 (t) i1 (t) L11 L12
=L· where L = (6.24)
λ2 (t) i2 (t) L21 L22
6-8 MODULE 6. ELECTRICAL SYSTEMS

is the inductance matrix.


According to Faraday’s law, voltages across the inductors
dλ1 dλ2
v1 (t) = + R1 i1 and v2 (t) = + R2 i2 (6.25)
dt dt
Hence, by differentiating (6.24), we can arrive at a constitutive relation of the coupled inductors in the
form of two differential equations
     
v1 (t) d i1 (t) i1 (t)
=L· +R (6.26)
v2 (t) dt i2 (t) i2 (t)
where  
R1 0
R= (6.27)
0 R2
By energy considerations it is established that in general
L21 = L12 = ±M (6.28)
where M ≥ 0 is the coupling factor. Hence, L is a symmetric matrix. Instead of the coupling factor M ,
mutual inductance is often expressed in terms of Also the coefficient of coupling
M
k= √
L11 L22
is often used to express mutual inductance. Note, that k ≤ 1. For k = 1, when M 2 = L11 L22 , the
coupling is termed ideal.
The sign of a mutual inductance depends on the orientation of the coil windings, and can be determined
by the dot-reference convention used also in Fig. 6.7. The dots denote either the ‘beginnings’ or the
‘ends’ of all the coil windings taking into consideration a unique orientation of the coil turns. The mutual
inductance of any two of coupled inductors is positive, if the variable reference of each of the two engaged
inductors is directed either towards, or from, its dot. If it is not so, the sign of the mutual inductance is
negative. In the particular example of the dot reference specification given in Fig. 6.7b, L21 = L12 ≤ 0.

6.7 More coupled inductors


If we have more than two linear time-invariant inductors coupled, (6.24) can be easily extended. For
example, if there are three coupled inductors like in Fig. 6.8, we have
     
v1 (t) i1 (t) L11 L12 L13
 v2 (t)  = L · d  i2 (t)  where L =  L21 L22 L23  (6.29)
dt
v3 (t) i3 (t) L31 L32 L33
In this case, L11 , L22 , L33 are the self-inductances, and
Ljk = Lkj = ±Mjk , k, j = 1, 2, 3, k 6= j
are the mutual inductances of the three coupled inductors. Taking into consideration the dot references
given in Fig. 6.8, L12 = L21 = −M12 , L23 = L32 = M23 and L13 = L31 = −M13 in this particular
example.
Regardless of the number of coupled inductors, (6.26) can be expressed as
d
v(t) = L · i(t) (6.30)
dt
or, in the integral form as Z t
i(t) = L−1 v(τ ) dτ + i(0+) (6.31)
0
where i(0+) is the vector of initial currents of the coils just right after t = 0.
Note that if the coupling between any of the involved inductors is ideal, the inverse inductance matrix
L−1 does not exist, because L is singular. This property follows from the fact, that conductances of any
two ideally coupled inductors – let us say between Lj and Lk – are interrelated, as M 2 = Ljk Lkj = Ljj Lkk .
Thus the corresponding 2 × 2 submatrix in L is singular.
6.8. COUPLED INDUCTORS AND IDEAL TRANSFORMER 6-9

Figure 6.8: (a) Three interacting coils, (b) coupled electrical inductors.

6.8 Coupled inductors and ideal transformer


Fig. 6.9 shows two models equivalent to the two coupled inductors in Fig. 6.7b. Both models are based
on ideal transformers. In the case of the model shown in Fig. 6.9a, where n = 1, the only purpose of the
unit-ratio transformer is to ‘isolate’ the ports corresponding to the inductors from each other.

Figure 6.9: Models equivalent to coupled electrical inductors employing an ideal transformer.

In the model given in Fig. 6.9b, the ideal-transformer ratio and the inductances are

L12 M2 M2
n= Lm = La = L11 − (6.32)
L22 L22 L22
Lm is called the magnetizing inductance as it models the effect of the magnetic flux common to both
coils. La is the leakage inductance modeling the effect of the leakage fluxes seen at the first port. The
equivalence of both models to a pair of coupled inductors can be easily verified by their analysis.
It is interesting to investigate the model in Fig. 6.9b for the limit case when L11 → ∞, L22 → ∞,
and k → 1. From (6.32) we can see that under these assumptions La → 0 and Lm → ∞. Hence, in the
limit case, the inductance La becomes a closed circuit while the inductance Lm becomes an open circuit.
Also, in the limit case, n = ±N2 /N1 assuming that N1 and N2 are the coil turns fully linked by their
fluxes. The sign of n can be determined in the same way as the sign of the mutual inductance.
6-10 MODULE 6. ELECTRICAL SYSTEMS
Module 7

Fluid systems

7.1 Fluid energy interactions


7.1.1 Fluid pressure
Pressure is the quantity which causes (or results from) fluid flow or sustains the weight of a column of
fluid. Fluid pressure is defined as the normal force exerted on a surface (real or imaginary) in a fluid
per unit area. When dealing with fluids contained in pipes or ducts, we shall assume pressure uniform
over the cross section areas of the pipes or ducts. Therefore, the total axial force F exerted on the cross
section area A is
F = pA
where p is the average bulk pressure over A.
Pressure is measured in the unit called a pascal, 1 Pa = 1 N/m2 . Two other commonly used pressure
units are the bar, 1 bar = 100 kPa, and standard atmosphere, 1 atm = 101.325 kPa. The absolute pres-
sure is measured relative to absolute vacuum. Most pressure-measuring devices, however, are calibrated
to read zero in the atmosphere, and so they indicate the difference between the absolute zero pressure
and the local atmospheric pressure. This difference is called the gauge pressure. Throughout this text,
the pressure p will denote gauge pressure unless it is otherwise specified.
Gauge pressure is considered positive if the fluid is in the state of compression. Since most fluids will
not support tension without cavitation (evaporation), we shall be concerned only with positive pressures
(cavitation requires a treatment utilizing thermodynamics).

Figure 7.1: Pressure across (a) a flow-through component, (b) an open tank supplied by liquid.

Fig. 7.1a shows a fluid-system component with two inlets. The bulk values of the gauge pressures at
the inlets A and B are denoted as pA and pB , respectively. The empty-head arrow in Fig. 7.1 indicates the
assumed positive polarity of the difference pBA = −pAB between the inlet gauge pressures. The pressure
difference
pBA = pB − pA (7.1)
is positive if pB > pA . Such a pressure difference is called the pressure drop across the component.
Another fluid-system component with two inlets is shown in Fig. 7.1b. It is an open tank in the form
of a cylinder of the cross sectional area A filled with a liquid coming to the tank via the inlet C. The
liquid is leaving the tank through the opening D situated in the tank wall close to its bottom. Recollect
that the pressure at any point in a fluid at rest is the same in all directions. If the fluid flow through the

7-1
7-2 MODULE 7. FLUID SYSTEMS

opening D is slow enough and the diameter of the opening is negligible when compared with the hight h
of the liquid in the tank above the tank bottom, we can assume that the pressure pD at the opening D
is the same as the pressure acting on the bottom of the tank. Thus,
mg
pD = (7.2)
A

where m is the mass of the liquid and g is the gravitational acceleration.


The mass m of a fluid enclosed in a volume V equals

m = ρV (7.3)

where ρ is the mass density of the fluid defined as the mass of the fluid per unit volume with units
kg/m3 .
Therefore,
V ρg
pD = = hρg (7.4)
A
as the volume of the liquid in the tank is V = A h.

7.1.2 Mass and volume flow


A liquid or gas usually flows between system components through pipes or ducts. In most practical
applications, such a fluid flow can be approximated to be one-dimensional flow. That is, the properties
can be assumed to vary in the direction of flow only. As a result, the properties such as pressure, density
or temperature are assumed to have uniform bulk average values over each cross section. They may
change, however, with time.
The one-dimensional mass-flow rate ṁ of fluid flowing in a pipe or duct is considered as

dm
ṁ = = ρ vA
dt

where A [m2 ] is the cross-sectional area of the pipe or duct normal to the flow direction, ρ is the average
mass density over A, and v is the average value of the fluid velocity in the flow direction over A.

Figure 7.2: Inlet volume flow of two-inlet components: (a) flow-through component, (b) open tank
supplied by liquid.

As measuring of volume-flow rate is much easier then measuring mass-flow rate, engineers prefer the
former variable when analyzing fluid systems. Volume-flow rate Q is the variable which expresses the
volume of fluid passing a given area per unit time, so its unit is m3 /s. This quantity, denoted in Fig. 7.2
by full-head arrows, is given as
dV
Q(t) = = vA (7.5)
dt
where V is the fluid volume and v is the bulk velocity of the fluid flow through the area. Note, that the
relation between the mass and volume flow rates is

ṁ = ρ Q
7.2. MODELING OF FLUID SYSTEMS 7-3

7.1.3 Fluid work, power and energy


If a quantity of fluid in a pipe or conduct is passing an area A in the direction of p, the amount of work
done by the force F = p A in moving this fluid through a distance dx is

dW = p A dx
The quantity of volume dV of fluid passing A is

dV = A dx

Therefore
p = dW/dV (7.6)
Thus pressure is the work done in passing a unit volume of fluid across the area on which p acts.
Power, as we know, is the rate of flow of work. Hence from (7.5) and (7.6), the fluid power

dW dV
P= =p = pQ
dt dt
The net power delivered to a two-inlet component shown in Fig. 7.1 is

P = pA Q A + pB Q B

If the fluid and the component is such that QA + QB = 0, i.e. if the flow in equals to flow out,

P = pAB QA = pBA QB

The fluid energy delivered to a fluid-system component is the time integral of the power.

7.2 Modeling of fluid systems


7.2.1 Fluid multipole diagrams
Modeling of fluid systems is very easy if the system can be represented by multipole models of system
components interconnected by ideal links. Fig. 7.3a shows the graphical symbol of a twopole model
representing a two-inlet component. The poles A and B corresponding to the component inlets are
denoted by short line segments called pins. The pins are associated with volume-fluid flows into the
components QA and QB as well as with the pole pressures pA and pB with respect to a reference (note
its symbol).

Figure 7.3: (a) Two-pole element. (b) Ideal fluid link.

The line segment shown in Fig. 7.3b represents a segment of an ideal pipe with inlets A and B
containing an ideal fluid such that

pA = pB and QA + QB = 0

Symbols of fluid-system components with their pins interconnected by ideal fluid links form a multipole
diagram of the system. The multipole diagram in Fig. 7.4a shows the way of measuring pressures of a two-
inlet component using ideal pressure-measuring instruments. The way of direct measuring of volume-flow
rate between elements is shown in Fig. 7.4b.
Note, that pressure is the across-variable in fluid systems. To measure the volume-flow rate between
components, we must sever the pipe interconnecting the components at the section of interest and include
7-4 MODULE 7. FLUID SYSTEMS

Figure 7.4: (a) Measurement of (a) element pressures and (b) of volume-flow rate between elements.

a flow-rate measuring instrument to renew the connection. It is thus evident that the volume-flow rate
is the through-variable.
Relation (7.1) is an example of the geometric constraints which are placed on pressures in a system
due to continuity of space. It is the expression of the postulate of pressure compatibility – called
also the axiom of pressure composition – for the two-inlet subsystem.
Mass, like energy, is a conserved property, and it cannot be created or destroyed. Mass m and energy
E can be converted to each other according to the famous formula proposed by Einstein: E = m c2 where
c = 2.998.108 m/s is the speed of light. This equation suggests that the mass of a subsystem may change
when its energy changes. However, for all energy interactions encountered in engineering practice, with
the exception of nuclear reactions, the change in mass is extremely small and cannot be detected by even
the most sensitive devices.
We can thus assume validity of the principle of conservation of mass according to which the mass
entering a region in space minus the mass leaving it must equal the change of mass stored within the
region.

Figure 7.5: Threepole model of a two-inlet component to satisfy the postulate of continuity.

Let us consider the component with two fluid inlets shown in Fig. 7.2. The volume flow rates at the
component inlets A and B are denoted as QA and QB . The assumed positive directions of the fluid flows
are indicated by the full-head arrows.
The principle of conservation of mass for the two-inlet component

ṁA + ṁB + ṁ = 0 (7.7)

where ṁA = ρ QA and ṁB = ρ QB is the mass-flow rate into the inlet A and B, respectively. As m = ρV
is the mass of the fluid contained in the component if V is the fluid volume of the component,

ṁ = ρ̇V + ρV̇ (7.8)

where ρ̇ = d ρ/dt and V̇ = dV /dt is the time rate of the mass density and component volume, respectively.
Unlike mass, fluid volume is not a conserved property as (7.8) indicates. Only if ρ̇ = 0, i.e., if the fluid
is incompressible, and if V̇ = 0, i.e. if the fluid volume in the component is constant, the flow entering
the component through one inlet equals the flow leaving the component through the other inlet

QA + QB = 0, or QA = −QB (7.9)

This is the postulate of continuity expressed for the fluid flow into the two-inlet component, which
is valid under the above assumptions only. If the fluid cannot be considered as incompressible, or the
container volume as constant, we must respect accumulation of fluid mass in the component by adding
a third pole C to the two-inlet component model linked to the reference as shown in Fig. 7.5. To satisfy
the postulate of continuity, the volume-flow rate of the third pole must be

QC = −QA − QB
7.3. FLUID ACCUMULATION 7-5

Fig. 7.6b shows three ideal pipe segments passing fluid freely between any two of the inlets A, B and
C via the interconnection D. Such an ideal-pipe structure is characterized by the relations

pA = pB = pC QA + QB + QC = 0

where the latter statement is the expression of the postulate of continuity for the interconnection C.

Figure 7.6: (a) Ideal pipe segment. (b) Ideal-pipe interconnection.

7.3 Fluid accumulation


7.3.1 Pure fluid capacitor
We shall define a pure fluid capacitor as an idealized physical element in which fluid volume V is a
function of pressure p inside the element
V = f (p) (7.10)
For an ideal fluid capacitor the relation (7.10) becomes linear, i.e.,

V (t) = Cf · p(t) (7.11)

where Cf [m3 .Pa−1 ] is the fluid capacitance of the capacitor.

Figure 7.7: (a)Symbol and (b) characteristics of pure fluid capacitor.

The relationship between the volume-flow rate of fluid into the capacitor and the fluid pressure at the
capacitor inlet is
dp
Q(t) = Cf · (7.12)
dt
or Z t
1
p(t) = Q dτ + p(t0 ) (7.13)
Cf t0
where p(t0 ) is the capacitor pressure at t = t0 .
Figure 7.7a gives our symbol for pure fluid capacitors. Only one the pole representing the fluid intake
of the capacitor is shown in the symbol. This pole is associated with the pressure p at the intake and
with the volume-flow rate Q(t) into the capacitor. The second pole is hidden as it must be always fixed
to the pressure reference.
The symbol consists of a square representing the capacitor volume and pole representing the capacitor
inlet. The capacitor is associated with two variables, the capacitor pressure p measured with respect to
a constant pressure reference, and the capacitor volume-flow rate Q. The assumed positive polarity of
p is indicated by the empty-head arrow placed between the symbol for the pressure reference and the
7-6 MODULE 7. FLUID SYSTEMS

symbol pole. The orientation of the full-head arrow shows that the positive Q is always directed into the
capacitor.
In general, a fluid capacitor can be nonlinear and time-variable. Then it is characterized by the
relation
dV ∂V dp ∂V
Q(t) = = · + (7.14)
dt ∂p dt ∂t
Figure 7.7b shows V -p characteristics of a nonlinear and linear fluid capacitors specified by their
constitutive relations (7.10) and (7.11).
Power P supplied to a capacitor is
P = Qp (7.15)
providing that the fluid density is constant.
Energy Ep , called fluid potential energy and stored in a fluid capacitor by virtue of the pressure, is
Z t
Ep = P dt (7.16)
0

dV = Cf dp (7.17)

p
Cf p2 V2
Z
Ep = Cf p dp = = (7.18)
0 2 2Cf

Open cylindrical tank


As a first example of a fluid capacitor we shall consider the open tank or reservoir of a cylindrical
shape shown in Fig. 7.8a. Let us assume that the liquid in the tank is incompressible, the tank walls are
rigid, and the flow resistance through the opening A at the bottom of the tank is negligible. Then the
volume of fluid in the tank is
V = Ah (7.19)
where A is the tank horizontal area and h is the height of the liquid column in the tank.

Figure 7.8: (a) Open cylindrical tank, (b) its fluid capacitor model.

The fluid pressure by the liquid level in the open tank is zero as it is the ambient reference pressure.
If there are no fluid accelerations in the tank except due to gravity, the fluid pressure by the tank bottom
equals to the weight of the liquid per the tank horizontal area, or
mg
p= = ρgh (7.20)
A
where m = ρ A h [kg] is the liquid mass, ρ [kg/m3 ] is the liquid mass density, and g [m/s2 ] is the
acceleration due to gravity.
Substituting p for h from (7.20) and comparing the result with (7.11) we obtain the open tank
capacitance as
A
Cf = (7.21)
ρg
Fig. 7.8b shows an ideal fluid capacitor of the fluid capacitance Cf modeling dynamics of the open
tank. In Fig. 7.9a, the liquid volume flow Q is supplied into the open tank by a pump. This situation
is modelled in Fig. 7.9b where the pump is represented by an independent source of fluid flow. The
relationship (7.20) between the height h of the liquid level and the pressure pA in the tank is expressed
in the model using a scaling block.
7.3. FLUID ACCUMULATION 7-7

Figure 7.9: (a) Open tank supplied by a pump, (b) model of the fluid system.

Open spherical tank


As an example of a nonlinear fluid capacitor let us consider a spherical tank with a fluid outlet at the
bottom and a vent at the top shown in Fig. 7.10a.

Figure 7.10: (a) Open spherical tank, (b) its V -p characteristic and (c) capacitance vs. liquid-level height.

The volume of liquid in the spherical tank is


π 2
V (h) = h (3r − h) (7.22)
3
and
dV
= π(2h r − h2 ) (7.23)
dh
where r is the tank radius and h is the height of the liquid level in the tank.
After expressing the pressure p by the tank outlet in terms of h using (7.20) we arrive at
 
π 1 2
Cf (p) = 2r p − p (7.24)
ρg ρg
Fig. 7.10b shows the constitutive relationship of the spherical tank, and the relationship Cf (h) is plotted
in Fig. 7.10c. Note that
4
Vmax = π, pmax = 2ρgr and Cf max = πr2
3

Rigid pressurized tank.


Another example of a component which may behave approximately like an ideal fluid capacitor is a closed
pressurized tankor chamber with rigid walls completely filled with fluid as shown in Fig. 7.11a. The
energy is stored in the tank because of the compressibility of the fluid.
In this case we shall neglect the possible variation of pressure in the tank due to gravity, and consider
only the variation in pressure which occurs as a result of the amount of fluid that has been forced into
the tank. Fluid inertia and frictional effects will be also neglected.
According to the principle of conservation of mass the time rate of mass entering the chamber through
the inlet A must equal the increase of mass in the chamber, or
d dρ
ρQA = (ρV0 ) = V0 (7.25)
dt dt
7-8 MODULE 7. FLUID SYSTEMS

Figure 7.11: (a) Tank containing compressed fluid. (b) Model of the pressurized tank.

where ρ [kg/m3 ] is the fluid density and V0 [m3 ] is the constant volume of the tank.
To employ tank pressure p rather than the mass density ρ as a variable, we must utilize the equation
of state of the fluid. For most liquids, a reasonably good first-order approximation for their equation of
state is
ρ
dρ = dp (7.26)
β
where β [N/m2 ] is the bulk modulus of elasticity of the liquid.
For most gases including air,
ρ
dρ = dp (7.27)
np
where p is the absolute fluid pressure and n is a constant whose value is between 1.0 and 1.4, depending on
how rapidly the fluid is compressed. If the pressure variations in the gas are small, it may be considered
to have a bulk modulus of elasticity β ≈ n p.
Substituting (7.26) into (7.25) yields
ρ dp
QA = (7.28)
β dt
Comparing this result with (7.14) indicates that the pressurized tank behaves like an ideal fluid capacitor
with the capacitance
V0
Cf = (7.29)
β
Under the above assumptions the pressurized chamber given in Fig. 7.11a can be represented graphi-
cally as shown in Fig. 7.11b. The capacitor pole representing the accumulator inlet is coalesced with node
A of a system model. The capacitor pressure p corresponds to the gauge pressure pA of the accumulator.

7.3.2 Equivalent bulk modulus


The bulk modulus of a liquid can be substantially lowered by gas or vapor entrained in the liquid in the
form of bubbles and/or pockets and by the mechanical compliance of a flexible container. The equation
which gives the equivalent bulk modulus of liquid-gas mixture in a flexible container is
1 1 1 Vg 1
= + + · (7.30)
βe βl βc Vt βg
where the subscripts g, l, and c refer to the gas, liquid, and container, respectively. It is assumed that the
initial total volume of the container Vt = Vl +Vg , and that βl >> βg . Thus the effective bulk modulus will
be less than any one of the values βl , βc , or Vt /Vg βg . The bulk modulus of the liquid βl is obtained from
manufactures’ data. The isothermal bulk modulus of air is equal to the pressure level p, the adiabatic
value used for air is (Cp /Cv )p = 1.4p.
Usually, the major source of mechanical compliance is the hydraulic lines connecting valves and pumps
to actuators. The bulk modulus for a thick-walled cylindrical container is
1 2 (1 + ν)D02 + (1 − ν)D2
= · (7.31)
βc E 2δ(D0 + D)
where D is the inner diameter, D0 is the outer diameter, δ is the wall thickness (2δ = D0 − D), E is the
modulus of elasticity of the wall material, and ν is the Poisson’s ratio for the material. For a thin-walled
cylinder such that D0 = D, and because ν = 0.25 for metals, (7.31) approximates to the formula
δE
βc = (7.32)
D
7.4. FLUID FLOW RESTRICTIONS 7-9

generally used for hydraulic tubing.


Estimates of entrapped air in hydraulic systems run as high as 20% when the fluid is at atmospheric
pressure. As pressure is increased, much of this air dissolves into the liquid and does not affect the bulk
modulus. In any practical case it is difficult to determine the effective bulk modulus other than by direct
measurement.

7.3.3 Temperature variations of liquids


The density of a liquid is a function of both pressure and temperature. A function relating density,
pressure, and temperature of a fluid is, by definition, the equation of state. The equation of state for
a liquid cannot be mathematically derived from physical principles. In contrast, the kinetic theory of
gases yields an equation of state. However, because changes in density as a function of pressure and
temperature are small for a liquid, the first three terms of a Taylor’s series for two variables may be used
as an approximation. Therefore,
   
∂ρ ∂ρ
ρ = ρ0 + (p − p0 ) + (ϑ − ϑ0 ) (7.33)
∂p ϑ ∂ϑ p

where ρ, p, and ϑ are the mass density, pressure, and temperature, respectively, of the liquid about initial
values of ρ0 , p0 , and ϑ0 . A more convenient form of (7.33) is
 
1
ρ = ρ0 1 + (p − p0 ) − α(ϑ − ϑ0 ) (7.34)
β
where    
∂p 1 ∂ρ
β ≡ ρ0 and α ≡ −
∂ρ ϑ ρ0 ∂ϑ p

Equation (7.34) is the linearized equation of state for a liquid. The mass density increases as pressure
is increased and decreases with temperature increase. Because mass density is mass divided by volume,
equivalent expressions for β and α are
   
∂p 1 ∂ρ
β = −V0 and α =
∂ρ ϑ V0 ∂ϑ p

where V is the total volume and V0 is the initial total volume of the liquid.
The quantity β is the change in pressure divided by the fractional change in volume at a constant
temperature and is called the isothermal bulk modulus or simply bulk modulus of the liquid. It is
the most important fluid property in determining the dynamic performance of hydraulic systems related
to the ‘stiffness’ of the liquid. Note that it is always a positive quantity. The reciprocal of β is often
termed the compressibility of the liquid.
An adiabatic bulk modulus, βa may also be defined. It can be shown that the adiabatic bulk
modulus is related to the isothermal bulk modulus by
Cp
βa = β (7.35)
Cv
where Cp /Cv is the ratio of specific heats. Because this ratio is only slightly in excess of unity for a
liquid, the distinction between adiabatic and isothermal moduli is insignificant.
The quantity α is the fractional change in volume due to a change in temperature and is called the
cubical expansion coefficient of the fluid.

7.4 Fluid flow restrictions


7.4.1 Pure fluid resistor
We shall now define a pure fluid conductor as a physical element characterized by the function

Q = f (p) (7.36)

where Q = 0 when p = 0, and the signs of Q and p are the same.


7-10 MODULE 7. FLUID SYSTEMS

For an ideal fluid conductor,


Q = Gf p (7.37)
where Q is the volume-flow rate of the conductor, p is the pressure drop across the conductor, and Gf
[m3 Pa−1 s−1 ] is the fluid conductance. The reciprocal value Rf = 1/Gf is the fluid resistance characterizes
the physical elements called fluid resistors.
It is important to note that (7.36) and (7.37) hold only when the rate of change of flow through the
conductor is small enough to avoid fluid inertia effects.
We see that fluid resistance dissipates fluid power. For incompressible fluids. the dissipated power P
is

P = p Q = Rf Q2 = p2 /Rf

Figure 7.12: (a) Pipe segment. (b) Pure fluid resistor and (c) conductor. (d) Pipe model.

Fig. 7.12b gives our graphical symbol for pure fluid resistors. The pure conductor is associated with
two variables: the conductor pressure drop p and the conductor fluid flow Q. Polarities of these two
variables are always related to the polarity of the symbol indicated by the + sign as shown in Fig. 7.12b.
This means that the pressure at the + pole is assumed larger than that at the second pole. The conductor
flow is considered positive if it is oriented from the + pole towards the conductor.
The symbol of the pure fluid conductor is utilized in Fig. 7.12c to model the pipe segment shown
in fig Fig. 7.12a. The nodes A and B represent energy interactions at the corresponding inlets of the
real pipe segment. The gauge pressures pA and pB at the pipe inlets are denoted by empty-head arrows
placed between the symbols for the reference pressure and the related nodes. Respecting the conductor
orientation, the pressure drop pBA corresponds to the conductor pressure p, and the flow QB corresponds
to the conductor flow Q. As the postulate of continuity applies to the pure conductor QA + QB = 0, or
QA = −QB .

Laminar flow through pipes


Fluid flows are generally dominated either by viscosity or inertia of the fluid. The dimensionless ratio of
inertia force to viscous force is called Reynolds number defined by
ρvd
Re = (7.38)
ν
where ρ is the fluid mass density, v is the average velocity of the flow, and ν [N.s.m−2 ] is the fluid
kinematic viscosity.
Flow dominated by viscosity forces is referred to as laminar or viscous flow. Laminar flow is
characterized by an orderly, smooth, parallel line motion of the fluid. Inertia dominated flow is generally
turbulent and characterized by irregular, erratic, eddy like paths of the fluid particles.
Reynolds number inside a pipe of circular cross section with diameter d is given by
4ρ Q
Re = (7.39)
πνd
where Q is the volume-flow rate, and µ is the absolute viscosity of the liquid. Transition from laminar to
turbulent flow has been experimentally observed to occur in the range 2000 < R < 4000. Below R = 2000
the flow is always laminar.
For example, the laminar incompressible flow of a liquid through a long capillary tube shown in
Fig. 7.12a can be approximated by the Hagen-Poiseuille law according to which the fluid conductance
of the capillary is
128νℓ
Rf = (7.40)
πρd4
7.4. FLUID FLOW RESTRICTIONS 7-11

where ν is the absolute viscosity of the fluid, ℓ is the length and d is the inside diameter of the capillary.
The ideal fluid resistor is applicable if the tube is long enough in relation to its diameter (ℓ ≫ 100d)
so that the effects of nonuniform flow near the tube entrance are negligible. At the same time, the rate
of change of volume flow must be small enough that inertia effects are negligible.
Also the effect of an incompressible flow through a porous plug can be approximated by an ideal
fluid conductor. In this case the constitutive relation (7.37) is expressed by the D’Arcy’s law. The
corresponding linear resistance Rf is usually determined for a given porous medium and working fluid
experimentally.
Example

Figure 7.13: (a) Discharging water tank. (b) Reservoir model.

Fig. 7.13a shows a water tank which discharges to the atmosphere through a porous plug. Recollect
that such a fluid restrictor produces a pressure drop linearly proportional to the flow through it. A
dynamic model of the tank is shown in Fig. 7.13b. Node A represents the interconnection of the plug
with the tank. Note that the direction of the volume-flow rate Q is in agreement with the direction with
the flow of the conductor but opposite to that of the capacitor.

Turbulent flow through pipes


When the flow rate through a tube becomes large so that the Reynolds number than 4000, the flow
through a tube is very likely to be turbulent, and the following approximate relation exists:
p = at Q |Q|3/4
where the bars indicate the absolute value and at is a constant whose value depends on the volume-flow
rate at which transition to turbulent flow occurs, the dimensions of the pipe (diameter and length), the
properties of the fluid (density ρ and viscosity µ), and the roughness of the inner walls of the pipe.
For a smooth pipe
0.3164 8ℓρ0.75 ν 0.25
at = 0.25 (7.41)
4 π 0.75 d4.75
Turbulent flow in closed conduits of non-circular cross section may be approximately computed from
(7.41) if the diameter is considered to be the hydraulic diameter defined by
4A
dh = (7.42)
S
where A is the flow section area and S is the flow section perimeter. The concept of hydraulic diameter
cannot be used for laminar flows because such flows are highly dependent on passage geometry. It is
often simplest to employ an experimental measurement to determine at for a given specific pipe.

Flow through narrow orifices


An orifice is a sudden restriction of short length in a flow passage and may have a fixed or variable area.
Orifices are a basic means for the control of fluid power.
Let us assume a rectangular orifice of width h and of length b (perpendicular to the plane of the
figure). Assuming turbulent flow, the flow Q through the orifice can be found as
r
2p
Q = hbCd (7.43)
ρ
where p is the pressure drop across the orifice, ρ is the specific mass of the liquid, and Cd is a constant
discharge coefficient. Experience shows that the theoretical value Cd = π/(π + 2) = 0.611 can be used
for all sharp-edged orifices, regardless of the particular geometry, if the flow is turbulent and A0 ≪ A1 .
7-12 MODULE 7. FLUID SYSTEMS

The turbulent character of the flow is ensured only at large enough Reynolds numbers. At smaller
values of Re the flow becomes laminar and the discharge coefficient varies with Re and depends on the
geometry of the orifice edges. It is common in practice, however, to use the turbulent flow equation (7.43)
regardless of the Reynolds number unless flows through very small openings are studied.

7.5 Fluid inertia


The physical element called pure fluid inductor or pure fluid inertor models the inertia effects encountered
in accelerating a fluid in a pipe or passage. An ideal fluid inductor is defined by the linear relation

r(t) = Lf · Q(t) (7.44)

or
d
p(t) = Lf Q(t) (7.45)
dt
where r(t) is the inductor pressure momentum, p(t) is the inductor pressure, Q(t) is the inductor volume
flow, and the constant parameter Lf [Pa.s2 m−3 ] is the fluid inductance of the fluid inductor, also termed
fluid inertance.

Figure 7.14: (a) Pipe segment. (b) Pure fluid inductor. (c) Pipe model.

To visualize the concept of fluid inductance, let us consider the unsteady frictionless flow of an
incompressible fluid in a nonaccelerating pipe segment of length ℓ. If the pipe has constant area cross-
section A and the velocity v of the fluid is uniform across any cross section of the pipe, we can say that
every fluid particle has the same velocity v and hence the same acceleration dv/dt. Then the force F
necessary to produce an acceleration dv/dt of the fluid mass in the pipe is

dv
F (t) = ApAB (t) = Aℓρ
dt
where pAB (t) is the pressure drop across the pipe segment. As ρ [kg.m−3 ] denotes the mass density of
the fluid, ρAℓ is the mass of all the fluid in the pipe segment. The volume flow QA (t) = Av(t), so that
the fluid inductance of the flow in the pipe is

Lf = ρ (7.46)
A
This relation holds only when the pipe is not being accelerated. If the pipe itself has an acceleration,
additional pressure difference effect between the inlets of the pipe segment will have to be taken into
account.
Fig. 7.14b gives our graphical symbol for the pure fluid inductor (or resistor). The pure inductor is
associated with two variables: the inductor pressure drop p and the conductor fluid flow Q. Polarities of
these two variables are always related to the polarity of the symbol indicated by the + sign as indicated
in Fig. 7.14b. This means that the pressure at the + pole is assumed larger than that at the second pole.
The inductor flow is considered positive if it is oriented from the + pole towards the inductor.
The symbol of the pure fluid inductor is utilized in Fig. 7.14c to model the pipe segment shown in
fig Fig. 7.14a. The nodes A and B represent energy interactions at the corresponding inlets of the real
pipe segment. The gauge pressures pA and pB at the pipe inlets are denoted by empty-head arrows
placed between the symbols for the reference pressure and the related nodes. Respecting the conductor
orientation, the pressure drop pAB corresponds to the conductor pressure p, and the flow QA corresponds
to the conductor flow Q. As the postulate of continuity applies to the pure inductor QA + QB = 0, or
QA = −QB .
7.6. FLUID TRANSDUCERS 7-13

In actual fluid piping, significant friction effects are often present along with the inertance effects, and
the inertance effect tends to predominate only when the rate of change of flow rate (fluid acceleration) is
relatively large. Since flow resistance in a pipe decreases more rapidly with increasing pipe area A than
does inertance, it is easier for inertance effects to overshadow resistance effects in pipes of large sizes.
However, when the rate of change of flow rate is large enough, significant inertance effects are sometimes
observed even in fine capillary tubes.

Figure 7.15: (a) Two-tank system. (b) Dynamic model.

A fluid-flow system consisting of two interconnected tanks is shown in Fig. 7.15a. The corresponding
model respecting fluid-flow inertia is in Fig. 7.15b.

7.6 Fluid transducers


7.6.1 Fluid-mechanical-fluid transformer
Fig. 7.16a shows the principle of a fluid transformer. The device consists of two chambers filled with
fluid separated by a common piston. The cross-sectional areas Aa and Ab on each side of the piston are
different. As the forces acting on both sides of the piston must be balanced

pA AA = pB AB

where pA and pB are presures exerted on the piston by the fluid in each of the chambers. To obtain a
constitutive relation characterizing the pure model of the device shown in Fig. 7.16b we must consider
this expression together with the energy-transfer relation taking on the form

pA Q A + pB Q B = 0

where QA and QB are the fluid volume flows into the chambers through the inlets A and B. Thus the
constitutive relation is      
pA (t) 0 n QA (t)
= ·
QB (t) −n 0 pB (t)

where the coupling ratio


pA QB Ab
n= =− =−
pB QA Aa
To respect such effects like fluid compressibility and inductance, piston friction and mass, fluid leakage,
etc., the energy-transfer transformer should be combined with the corresponding physical elements.

Figure 7.16: Fluid transformer.


7-14 MODULE 7. FLUID SYSTEMS

7.6.2 Mechanical-fluid-mechanical transformer


Fig. 7.17a shows another fluid-based transformer, This time, however, the transformer is arranged to
convert rectilinear motions. It consists of a cylinder with two pistons the cross-sectional areas Aa and
Ab of which are different. The space between the pistons is filled with fluid exerting on both pistons the
same pressure p which is balanced by the external forces FA and FB acting in the pistons, i.e.,

p = FA /Aa = −FB /Ab

Combining this expression with the energy-transfer relation in the form

ẋA FA + ẋB FB = 0

we obtain the constitutive relation of the rectilinear-motion transformer pure model shown in Fig. 7.17b
     
ẋA (t) 0 n FA (t)
= ·
FB (t) −n 0 ẋb (t)

where
ẋA FB Ab
n= =− =
ẋB FA Aa

Figure 7.17: Fluid-based rectilinear-motion transformer.

7.6.3 Fluid-mechanical transducers


Table 7.1 shows pure models of some typical fluid-mechanical transducers utilizing energy-transfer gyra-
tors.

Figure 7.18: Three-land-three-way spool valves with cylinder load (S ... supply, R ... return).
7.6. FLUID TRANSDUCERS 7-15

Table 7.1: Pure models of fluid-mechanical transducers


Transducer Pure model

Figure 7.19: Spool valve model.


7-16 MODULE 7. FLUID SYSTEMS

Figure 7.20: Flow forces acting on spool valves.

Figure 7.21: Flapper valve.


7.6. FLUID TRANSDUCERS 7-17

Figure 7.22: Fluid-damped car suspension.


7-18 MODULE 7. FLUID SYSTEMS
Module 8

Thermal systems

8.1 Thermal interactions


8.1.1 Temperature
Temperature is the quantity related to the mean kinetic energy of molecules in a substance. When
the temperature of a substance varies, its various physical characteristics change. For example, as the
temperature of a metal increases, its dimensions and electrical resistance usually also increases, and its
color may change.
Such changes may be used in thermometers to establish a quantitative measure of temperature.
For instance, in a mercury-bulb thermometer, the scale of temperature is associated with the length
of the mercury column. The change in length of the column which occurs when the thermometer is
immersed first in ice water and then in boiling water is divided into equal increments called degrees.
In the Celsius or centigrade system [◦ C] the interval is divided into 100 one-degree increments, and the
ice-water temperature is defined as 0◦ C.
The absolute scale of temperature is independent of physical characteristics of substances. The second
law of thermodynamics shows that negative absolute temperatures are impossible. When the absolute
temperature is measured in degrees centigrade the scale is called the Kelvin scale and the units are kelvins
[K]. The freezing point of water 0◦ C is expressed as 273.2 K. We shall use the symbol ϑ to represent
temperature measured by any of the above scales. Thus the interrelation between the Kelvin and Celsius
scales can be expressed as
ϑ [K] = ϑ [◦ C] + 273.15
Consideration of the methods by which temperature could be measured reveals that temperature is
an across-variable (it can be measured without breaking into the system).

8.1.2 Heat flow


If two bodies having different temperatures are brought into contact, their temperatures will approach
each other in such a way that one temperature will increase and the other will decrease. The quantity
that has been transferred from one body to the other to produce the observed temperature changes is
called heat. If the body temperature is decreased by the transfer, the heat flow is considered to be out
of the body. Since temperature is a measure of the mean energy of the molecules of the body, heat is a
form of energy.
Heat can be measured in calories [cal]. One calorie is the amount of heat required to raise the
temperature of 1 gram of water by 1 ◦ C. Since heat is a form of energy, it may also be expressed in joules
by means of the equivalence of mechanical work and heat determined by the first law of thermodynamics.
Thus 1 cal is equivalent to 4.1868 J.
We shall use the symbol H for quantity of heat. The flow rate of heat Ḣ is obtained by differentiating
H
dH
Ḣ = (8.1)
dt
In the systems we have studied previously, power has been given by the product of the through-
and across variables at any energy interaction in the system. In thermal systems this situation does not
prevail, since heat flow – the through-variable – is itself the thermal power.

8-1
8-2 MODULE 8. THERMAL SYSTEMS

8.1.3 Thermal power and energy


Thermal power is thus
P = Ḣ (8.2)
Thermal power is measured in watts [W].
The thermal energy E transferred through a surface (real or imaginary) in a body is the time integral
of the heat flow Z t
E= Ḣ dτ
0

8.2 Composition of thermal interactions


8.2.1 Composition of temperatures
The temperature drop between two points A and B in a system equals the difference between temperatures
of these points, i.e.
ϑAB = ϑA − ϑB (8.3)
This is the zeroth law of thermodynamics generalized for transient states of thermal systems. The
zeroth law states that if two bodies are in thermal equilibrium with a third body, they are also in thermal
equilibrium with each other. Bodies in contact are said to have reached thermal equilibrium if the heat
transfer between them stops.

8.2.2 Composition of heat flows


The first law of thermodynamics tells us that for any subsystem defined by a closed imaginary surface,
the total heat into the subsystem must equal the energy transferred out of the subsystem in other forms
(such as mechanical work, electrical energy, or fluid energy) plus the increase of energy stored in the
system.

8.3 Accumulation of heat


8.3.1 Pure thermal capacitors
The ability of a body to store internal energy by virtue of a temperature rise when it receives a flow of
heat can be modelled by a pure thermal capacitor. Let us consider heat flow taking place into a body
having uniform (but not necessarily constant) temperature ϑ relative to a constant reference temperature.
By definition, a pure thermal capacitor modeling the thermal-energy-storage phenomenon has a
heat content H which is a function of the body temperature ϑ

H = f (ϑ) (8.4)

In an ideal thermal capacitor, H and ϑ are proportional:

H = Ct ϑ (8.5)

where Ct is the thermal capacitance in J/K.


The thermal capacitance Ct of a body is determined as

Ct = cp m (8.6)

where m [kg] is the body mass, cp [J/kg] is the specific heat of the body material.
The symbol for a pure thermal capacitor is shown in Fig. 8.1a . As for a fluid capacitor and inertor,
one pole of the thermal capacitor must be always associated with a reference temperature which is zero
or constant. As no heat occurs through the reference “connection” the reference is not shown in the
symbol.
For a body to be modelled by an ideal thermal capacitor, all thermal resistance effects (discussed
below) should be negligible within it (usually in comparison with external resistances) so that the tem-
perature will be uniform throughout the mass which comprises the capacitor. A mass of metal or a
well-stirred tank of fluid often behave very nearly like ideal thermal capacitors.
8.4. HEAT TRANSFER 8-3

Figure 8.1: Symbol of pure thermal (a) capacitor, (b) resistor and (c) conductor. (d) Symbol of thermal
reference.

8.3.2 Energy of thermal capacitors


The energy E stored in a thermal capacitor in the absence of work effects is simply equal to the heat
transferred into it. Here again, the energy does not enter the picture in the same way as in the case of
elements in the other energy domains, because the stored energy is a function of only the first power of
the across-variable ϑ as
E = H = f (ϑ) (8.7)

8.4 Heat transfer


8.4.1 Pure thermal resistors
Heat can be transferred in three different ways: conduction, convection, and radiation. All modes of heat
transfer require the existence of a temperature difference, and all modes of heat transfer are from the
high temperature medium to a lower temperature one.
By analogy with the other pure resistors, we will define a pure thermal resistor by the equation

ϑ = f (Ḣ) (8.8)

An ideal thermal resistor has the defining equation

ϑ = Rt Ḣ (8.9)

where Rt [K/W] is the thermal resistance.


An ideal thermal conductor is defined as

Ḣ = Gt ϑ (8.10)

where Gt [W/K] is the thermal conductance.


All materials offer some resistance to heat flow, which is evidenced by the temperature drops in
the direction of heat flow through material. When the material offers a large degree of resistance to
heat flow, it is called a thermal insulator. A material through which heat flows relatively freely is
called a thermal conductor. (At very low temperatures near the absolute zero certain materials exhibit
“superconductivity”, in which case heat may flow practically at zero temperature difference.)

8.4.2 Thermal conduction


The flow of heat through a solid material, such as the sheet illustrated in Fig. 8.2, occurs by conduction,
that is, by the transfer of thermal kinetic energy from atom to atom of the material. This type of heat
flow is described quantitatively by Fourier’s law
σc A
Ḣ = ϑAB (8.11)

where A is the area normal to the heat flow direction, ℓ is the length in the flow direction, and ϑAB is the
temperature drop across the sheet. The thermal conductivity σc , expressed in W.m−1 .K−1 , is a specific
property of a material defined by (8.11).
Equation (8.11) relates the through-variable Ḣ to the across-variable ϑ by a simple constant factor.
Comparison of (8.11) and (8.9) shows that the thermal resistance of conduction is

Rt = (8.12)
σc A
8-4 MODULE 8. THERMAL SYSTEMS

Figure 8.2: Heat flow through a conducting material.

A given material may be considered as a pure or ideal thermal resistance only if the amount of heat
stored in it is negligible or does not change appreciably during operation. Fig. 8.1b shows the symbol for
pure thermal resistances.
When two sheets of solid material are placed together with imperfect contact (which is typical), the
total resistance of the combination is greater than the sum of the individual resistances of each of the
sheets because of the small air space (or vacuum or oxide film in some cases) which exists at the interface.
Although the thickness of the air space is very small, this extra resistance can be appreciable, because
still air is a relatively poor conductor of heat.

8.4.3 Thermal convection


When the material through which heat is transferred is a fluid (liquid or gas), one may compute its
resistance in much the same way as for a solid, but only so long as the fluid is not in motion. When fluid
motion occurs (natural convection, forced flow, etc.), the heat-transfer picture becomes very complex.
This heat-transfer mechanism is called convection, and is illustrated in Fig. 8.3. Motion, especially
if it is turbulent, reduces the resistance to heat flow through a fluid so much, that the major thermal
resistance exists in a very thin layer near the wall – described in terms of a film – where fluid motion is
small.

Figure 8.3: Convective heat transfere.

The notion of an over-all coefficient of heat transfer for a given physical arrangement of components
in a thermal system is frequently employed, and it is defined as

ch = (8.13)

where ch is the coefficient of heat transfer [kg.cal/oC.s.m2 ], A is the cross-sectional area, and ϑ is the
over-all temperature drop.
A determination of the heat-transfer coefficient for a given thermal situation is usually very complex
and nearly always involves experimental measurements. The coefficient ch is not constant for a given
geometry, but may often be assumed constant over limited ranges.
Equation (8.13) describes an ideal thermal resistance Rt whose value is
1
Rt = (8.14)
ch A

8.4.4 Thermal radiation


In addition to thermal conduction and convection, heat may be transferred to or from a body by thermal
radiation. Radiation is described quantitatively by the Stefan-Boltzmann law,
Ḣ = cr (ϑ4A − ϑ4B ) (8.15)
8.5. THERMAL INDUCTORS AND TRANSFORMERS 8-5

where cr is a constant which depends on the shape, area, and surface characteristics of the body which is
emanating or receiving radiation, and on the wavelength of the radiation. The temperatures ϑA and ϑB
must be the absolute temperatures in K of the surfaces emitting and receiving radiation, respectively.
Fig. 8.4, illustrates schematically radiative heat transfer. Here it is assumed that the radiant energy
emitted by surface A is absorbed by surface B.

Figure 8.4: Convective heat transfere.

The constant cr in (8.15), measured in W/K4 is very small, so that heat transfer by radiation is usually
insignificant unless large temperatures are involved. As an example, for radiation from a spherical black
body to a black environment which completely surrounds it, the value of cr is given by σAo , where σ is
the Stefan-Boltzmann constant, σ = 1.37 × 10−5 [W.m−2 K−4 ] and Ao is the area of the radiating sphere
in m2 .
The expression (8.15) for heat transfer is nonlinear and does not represent a pure thermal resistance
as we have defined it, because Ḣ does not depend on temperature difference. If the temperatures ϑA and
ϑB change by only small amounts from mean values ϑA0 and ϑB0 , (8.15) may be linearized.
Let Ḣ0 be the heat flow when ϑA = ϑA0 and ϑB = ϑB0 , let ϑB be constant and ∆ϑA be a small
change (compared with ϑA0 and ϑB0 ) of temperature ϑA . Then ∆Ḣ is the change in heat flow

∆Ḣ = 4cr ϑ3A0 ∆ϑA = ∆ϑA /Req


(8.16)
Ḣ0 = cr (ϑ4A0 − ϑ4B0 )

and Req is the equivalent linear resistance when Ḣ = Ḣ0 .


When a thermal-resistance effect occurs due to combined radiative and convective heat transfer, it is
often true that the over-all resistance is nearly linear and can be represented over the range of interest
by an ideal thermal resistance.
A thermal resistance, unlike mechanical, electrical, and fluid resistances, dissipates no energy since the
net heat flow is always zero. It also stores no energy since no work is done during the heat-flow process in
a thermal resistance (first law of thermodynamics). There is, however, an increase of the property called
entropy associated with the flow of heat through a thermal resistance (second law of thermodynamics),
and hence a loss or dissipation of the available part of the energy supplied to it at temperature ϑA . If
the heat flow Ḣ were put into a reversible heat engine (a thermal-to-mechanical transducer) an amount
of mechanical power equal to Ḣ(ϑA − ϑB )/ϑA could have been taken from the engine.

8.5 Thermal inductors and transformers


By analogy to electrical or fluid domain, there should be a pure thermal inductor as the second energy-
storage element. While the thermal capacitor stores energy by virtue of its temperature, the thermal
inductor should store energy by virtue of a heat flow through it. There is, however, no known thermal
phenomenon at ordinary temperatures which acts in this way. Thus there is no thermal element which
could be called a thermal inductor.
Pure transformer is a device which provides changes in the magnitudes of the through- and across-
variables while maintaining instantaneous power balance. Such a device is clearly impossible in thermal
systems because the through-variable itself is power (heat flow). Instantaneous power balance requires
that Ḣin = Ḣout , and therefore no transformation of Ḣ can occur.

8.6 Thermoelectric transducers


There are several kinds of thermal-to-electrical transducers. The most common of these is the thermo-
couple, in which a voltage difference is generated between two dissimilar metals when a temperature
8-6 MODULE 8. THERMAL SYSTEMS

difference is applied. This effect has been used for temperature measurement and for the generation of
small amounts of electric power. When energy flows from the electrical system to the thermal system, re-
frigeration or cooling can be obtained. Small thermoelectric refrigerators have been made for the cooling
of electric components and for use as automatically controlled constant-temperature references.
Module 9

Magnetic systems

9.1 Magnetic interactions


Magnetic field can be represented by lines of magnetic flux. Where these flux lines bypass the required
path through the circuit, they are termed flux leakage. Magnetic flux fringing is somewhat similar
to leakage and is a term used to describe the spreading of flux lines in an air gap of a magnetic circuit
thus increasing its effective area. The toroidal coil in Fig. 9.1a with windings of evenly distributed turns
forms a completely closed homogenous magnetic circuit without any leakage flux. This is true even if the
core is made from a nonmagnetic material. If the core of such a coil is made from a material with large
permeability, there will be no considerable leakage fluxes even if the coil winding is concentrated into one
place on the toroid, as indicated by the dashed line in Fig. 9.1a.

Figure 9.1: Magnetic field of different coils.

If the core is disrupted by an air gap, as in Fig. 9.1b, the flux lines indicate that large leakage and
fringing fluxes result and the flux within the core varies with the distance from the coil. Still, such a coil
can be modeled as a magnetic circuit if the largest part of the flux lines is concentrated into the core
which determines their geometric shape. In the contrary, the short cylindrical coil shown in Fig. 9.1c
cannot be modeled as a magnetic circuit. Except for the short part of the flux lines within the core, most
of their length is situated in the air.
Magnetic voltage (called also magnetomotive force) w(t) [A] plays the role of the across variable in
magnetic interactions. The concept of magnetic voltage is applicable in quasi-magnetic fields when the
rate of change of electric field normal to the subsystem energy boundary is negligible.
Magnetic flux rate ψ̇(t) [Wb/s] into a subsystem through an energy entry is defined as

φ̇(t) =
dt
where φ(t) is the total of magnetic flux B [Wb/m2 ] over the entry area passing through it in the perpen-
dicular direction to it.
Magnetic interactions are also governed by laws similar to those of Kirchhoff. The continuity postulate
applies to the rate of magnetic flux into the entries of a subsystem assuming the magnetic flux leakage
through the subsystem boundary is negligible. If it is not so, the boundary must be enlarged to include
also the leakage flux lines.

9-1
9-2 MODULE 9. MAGNETIC SYSTEMS

Figure 9.2: Postulate of continuity for magnetic circuits.

Fig. 9.2a shows branching of the magnetic flux in a ferromagnetic material. The dynamic diagram
representation of the branching is in Fig. 9.2b. The interconnection portrayed there by node A with three
line segments representing ideal magnetic conductors.
Obviously, the magnetic flux in magnetic systems obeys the continuity postulate based on the law of
conservation of magnetic field. In general, if there are n magnetic conductors interconnected at a point,
and if φk is a magnetic flux of the k-th conductor, then
n
X
φk = 0 (9.1)
k=1

where the fluxes oriented towards the interconnection are considered as positive.
According to Ampere’s law the integral around any closed path ℓ in the magnetic field of intensity H
equals the sum of electrical current I passing through the path ℓ, i.e.
I
H · dℓ = I (9.2)

Positive current is defined here as flowing in the direction of the advance of a right-handed screw turned
in the direction in which the closed path is traversed by H.
Assuming that the closed path ℓ is divided into n segments ℓk , k = 1, 2, . . . , n, then
I Xn Z
H · dℓ = Hℓ dℓ (9.3)
ℓ k=1 ℓk

where Z
Hℓ dℓ = wk (9.4)
ℓk

represents the magnetic voltage across the k-th segment of ℓ.


Hence, the integral (9.2) can be replaced by the sum
n
X
wk + wm = 0 (9.5)
k=1

where wm = −I represents the total magneto-motoric voltage of ℓ. We may thus reformulate Ampere’s
law as stating that “the sum of magnetic voltage drops around any closed path is equal to zero,” which
represents the compatibility postulate for magnetic systems.

9.2 Pure magnetic capacitor


A pure magnetic capacitor can be described by the constitutive relation

f (φ, w, t) = 0 (9.6)

For an ideal magnetic capacitor (9.6) becomes

φ(t) = Cm · w(t) (9.7)

where Cm , the magnetic capacitance, is known as the permeance, and is measured in henrys. Its reciprocal
value Dm = 1/Cm is called the reluctance.
9.2. PURE MAGNETIC CAPACITOR 9-3

The energy stored in the magnetic capacitor is


1 1 1 2
Z Z
W = φ̇w dt = Cm w dw = Cm w2 = φw = 2
φ (9.8)
2 2 2Cm
Magnetic flux φ through a part of a surface of an area S [m2 ] is defined by the integral
Z
φ= Bn dS (9.9)
S

where Bn is the magnetic flux density vector which is normal to the surface. Within a material, B and
H are related by the permeability µ of the material:
B = µH µ = µ0 µr (9.10)
where µ0 = 4π × 10−7 [H.m−1 ] is the permeability of free space, and µr [ – ] is the relative permeability
of the material. Permeability can be qualified as a scalar only in regions of homogenous and isotropic
materials. In case of nonmagnetic materials µr can be considered as equal to 1.0 for most of the practical
purposes. Properties of most of magnetic materials are strongly nonlinear, the expression (9.10) can then
be applied to small variations of H in the vicinity of a certain fixed operating point.

Figure 9.3: (a) Ferromagnetic core. (b) Symbol of pure magnetic capacitor. (c) Dynamic model.

Fig. 9.3a shows a piece of a core made of ferromagnetic material of a cross section area A and of
length ℓ. If we assume that the magnetic field intensity H is nonvarying along the length ℓ, and that the
flux density B is uniform over the area A, the permeance of the component becomes
φ(t) BA µA
Cm = = = (9.11)
w(t) Hℓ ℓ
which is similar to the expression for electrical capacitance in a region with similarly uniform electrical
properties.

Figure 9.4: (a) Ferromagnetic torus. (b) Dynamic model.

Fig. 9.4a shows a torus made of highly permeable material with a winding of N turns. The magnetic
voltage around the magnetic path is w = N i. This magnetic voltage establishes a magnetic field intensity
H, which in turn produces a magnetic flux density B = µ H in the material. The narrow air gap
interrupting the closed magnetic circuit makes an example of a magnetic system components that can be
modeled as an ideal magnetic capacitor as shown in Fig. 9.4b.
Magnetic circuit with leakage fluxes forms a distributed-parameter system. As in the case of other
similar systems, it can be approximated by decomposing it into sections each of which is represented
by a lumped parameter model. The larger is the number of the sections, the more accurate is the
approximation. For example, the magnetic circuit in Fig. 9.5a is divided into a number of parts and
assumed fluxes both in the core and in the air are indicated. Fig. 9.5b shows the corresponding lumped-
parameter model in which the permeances are related to the core segments, to the leakages and to the
gap.
9-4 MODULE 9. MAGNETIC SYSTEMS

Figure 9.5: Magnetic circuit with distributed parameters and its models.

9.3 Pure magnetic conductor


The constitutive relation of a pure magnetic conductor or resistor is

f (φ̇, w, t) = 0 (9.12)

In case of an ideal magnetic conductor (9.12) becomes

φ̇(t) = Gm · w(t) (9.13)

where Gm , measured in ohms, is the magnetic conductance of the conductor. The ideal magnetic resistor
is an element characterized by the inverse relation

w(t) = Rm · φ̇(t) (9.14)

where Rm is the magnetic resistance measured in siemens. Such elements can be used for modeling eddy
current losses in magnetic systems, for example. Ideal magnetic conductors or resistors can be used to
model energy dissipation losses associated with small variations of magnetic field in some materials.
Note, that for the sake of consistency in the unified approach to modeling of dynamic effects in different
energy domains, we do not adopt the traditional definitions of magnetic conductance and resistance. That
is, we do not consider magnetic conductance to be equivalent to permeance, and magnetic resistance to
reluctance as it is common on most of the other sources.

9.4 Ideal magnetic inductors


There is no known magnetic phenomenon at ordinary temperatures which acts as a magnetic inductor.

9.5 Magneto-electric conversion


Let us consider Fig. 9.6a with a coil consisting of N turns of wire wrapped around a ferromagnetic core
with an air gap. The coil winding is supplied by the electrical current i making an electrical voltage drop
v across the winding. The magnetic flux φ invoked in the core by the magnetic voltage N i is divided
by the two symmetrical core branches into two equal parts. The directions of the electrical current and
magnetic flux comply with the right-handed screw rule.
A magnetic model of the coil is given in Fig. 9.6b. The magnetic capacitor with the reluctance
Dg represent there the magnetic field energy accumulation in the air gap. Similarly, the capacitors of
9.5. MAGNETO-ELECTRIC CONVERSION 9-5

Figure 9.6: (a) Electromagnetic coil, (b) magnetic model, (c) simplified magnetic model.

reluctances Df 1 , Df 2 and Df 3 represent this accumulation in the individual branches of the ferromagnetic
core. In Fig. 9.6c, the same model is simplified taking advantage of the core symmetry.
Analyzing the model in Fig. 9.6c we shall obtain

Ni
φ= (9.15)
Dm

where Dm = Dg + Df 1 /2 + Df 3 is the coil reluctance.


So far we were discussing multipole models of the coil from the point of view of the magnetic field only.
To be able to interconnect the coil model with a model of an electrical circuit interacting with the coil,
we have to add a model of the electromagnetic conversion. In Fig. 9.7a, this conversion is represented by
the gyrator of the gyrating ratio equal to the number of the coil turns N . The gyrator model consists
of two controlled sources, both with the transfer ratio equal to N . One of them is an electrical voltage
source controlled by the magnetic flux rate, the second is a magnetic voltage source controlled by the
electrical current.

Figure 9.7: (a), (b) Coil electromagnetic models, (c), (d) coil electrical models.

Fig. 9.7b differs from Fig. 9.7a by elements modeling some parasitic effects in the coil. The electrical
resistor R corresponds to the electrical resistance of the coil winding, the magnetic resistors Rf 1 and
9-6 MODULE 9. MAGNETIC SYSTEMS

Rf 2 stand for the energy losses in the segments of the ferromagnetic core, and the electrical capacitor C0
represents accumulation of electric energy within the coil winding.
Note, that the model in Fig. 9.7b cannot be described just by a set of algebraic equations as it was in
the case of the other coil models discussed so far – here we need a set of differential equations.
Analyzing the model shown in Fig. 9.7c we can arrive at the coil electrical model given in Fig. 9.7c,
i.e. in a pure inductor. According to the static definition of the coil inductance

λ
L= (9.16)
i
where λ = N φ is the flux linkage measured in webers. Taking into consideration (9.15), the coil inductance
can be expressed as
N2
L= (9.17)
Dm
Similarly, analyzing the model from Fig. 9.7b we can obtain parameters of the coil electrical model
with parasitic effect elements in Fig. 9.7d.

9.6 Magnetic nonlinearities


In ferromagnetic materials, the material is divided up into small regions or domains in each of which
all dipole moments are spontaneously aligned. When the material is completely demagnetized, these
domains have random orientation resulting in zero net flux density in any finite sample. As an external
magnetizing force H is applied to the material, those domains that happen to be in line with the direction
of H tend to grow, increasing B. The resulting B-H magnetization characteristic, known also as the B-H
virgin curve, is shown in Fig. 9.8.
As H is further increased, this process continues until all domains are aligned with H. This results
in magnetic saturation and the flux density within the material cannot increase beyond the saturation
density Bs .

Figure 9.8: Magnetization characteristic.

The small increase that occurs beyond this condition is due to the increase of the flux density in the
vicinity of the space occupied by the material. It is often convenient to subtract out this component of
‘free space’ flux density and observe only the flux density variation within the material. Such a curve
is known as the intrinsic magnetization characteristic and is of use in the design of permanent magnet
devices.
A symmetrical hysteresis loop is obtained after a magnetic field intensity reversal between plus and
minus Hs . The area within the hysteresis loop corresponds to the hysteresis losses, i.e. to the energy
dissipated in the material during one reversal cycle. This area varies with temperature of the material
and with the frequency of the reversal.
Eddy current losses are caused by electrical currents induced in the core, called eddies, since they
tend to flow in closed path within the magnetic material itself. The eddy current loss in a sinusoidally
9.7. PERMANENT MAGNET EXCITED SYSTEMS 9-7

excited material, neglecting saturation, can be expressed by the relationship

Pe = ke f 2 Bm
2
[W] (9.18)

where Bm is the maximum value of flux density, f the frequency, and ke is a proportionality constant
depending upon the type of material and the lamination thickness.
To reduce eddy current loss, the magnetic material is laminated, that is, divided up into thin sheets
with a very thin layer of electrical insulation between each sheet. The sheets must be oriented in a
direction parallel with the flow of magnetic flux. It should be recognized that laminating a magnetic
material generally increases its volume. The ratio of the volume actually occupied by magnetic material
to total volume of a magnetic part is known as the stacking factor. This factor is important in accurately
calculating flux densities in magnetic parts.
Therefore, instead of Eq. (9.10), for larger variations of magnetic field intensity H the relation

B = f (H) (9.19)

expressing the flux density B as a function of H has to be considered.

9.7 Permanent magnet excited systems


The second type of excitation source that is commonly used for supplying energy to magnetic circuits used
in rotating machines and other types of electromechanical devices is the permanent magnet. The excita-
tion can be described with the aid of the second quadrant B-H curve, often called the demagnetization
curve.
Theoretically, it is possible to have infinite permeance in the external magnetic circuit that would
correspond to α = 90◦ and the magnet operating point would be at B = Br , H = 0. This situation
is approximated by a permanent magnet having an external circuit consisting of no air gap and a high
permeability member often called a ‘keeper’. In practice, however, there is always a small equivalent air
gap and a small reluctance drop in the keeper, and the operating point is to the left of Br .
For a finite air gap, the operating point will be at some point A on the B-H curve, and the permanent
magnet will develop the magnetic field intensity HA to overcome the reluctance drop of the air gap and
other portions of the external magnetic circuit. If the air gap is increased, Pge decreases and the magnet
must develop a larger magnetic field intensity Hd . If the air gap is subsequently returned to its original
value, the operating point will not return to A but, rather to C.
The slope of the loop is known as recoil permeability. Since it is a slope on the B-H plane, it is also
sometimes called incremental permeability. Recoil permeability is an important parameter of a permanent
magnet for applications with varying air gaps. In general, a permanent magnet is used most efficiently
when operated at conditions of B and H that result in the maximum energy product.
9-8 MODULE 9. MAGNETIC SYSTEMS
Module 10

Electromechanical systems

10.1 Conductor moving in magnetic field


The principle of a variety of electromechanical transducers is shown in Fig. 10.1a. A conductor of length
ℓ is placed there in parallel with the y-axis in a homogenous magnetic field with flux density B oriented in
the x-direction. If current i flows through the conductor in the y-direction, a z-directional external force
F is required to hold the wire in equilibrium against the magnetic force acting on it. The magnitude of
the force is given by the Lorentz force law
F = −B ℓ i (10.1)
The direction of the force is determined by the left-hand Fleming’s rule. If the conductor moves at a
speed ż through the magnetic field in the z-direction, a voltage v is generated in the conductor. According
Faraday’s law of magnetic induction the magnitude of the voltage is

v = B ℓ ż (10.2)

The polarity of v is given by Fleming’s right-hand rule. That is, if conductor moves in the direction of
the force F , the polarity of v with respect to i is such that an electrical power is generated.

Figure 10.1: (a) Current-carrying conductor moving in a magnetic field, (b) its ideal-transformer model.

Equations (10.1) and (10.2) form together the constitutive relationship of an ideal energy-transfer
transformer which can be used for modeling this simple device as illustrated in Fig. 10.1b. The transformer
poles A and B represent electrical terminals of the wire, whereas pole Cz represents motion of the wire
along the z axis. Coupling ratio of the transformer is obviously
ż 1 1
n= = = (10.3)
v Bℓ B2πrN

10.2 Coil moving in magnetic field.


As another example of an energy-transfer transducer, we consider a coil which can move in a uniform
magnetic field. The voice coil of a loudspeaker and the meter movement of an ammeter are representative
applications. Fig. 10.2a indicates a typical geometric configuration for a loudspeaker voice coil. A
permanent magnet maintains a uniform radial magnetic field B in the annular gap. Within the gap is
the voice coil, which can move longitudally with velocity ẋ. Flexible lead-in wires introduce a current i

10-1
10-2 MODULE 10. ELECTROMECHANICAL SYSTEMS

into the moving coil. The external force F acting on the coil is required to hold the coil in equilibrium
against the magnetic forces. The voltage across the voice coil is v.
The constitutive relations for the moving coil can be obtained from (10.1) and (10.2) by substituting
ℓ = 2π r N , where r is the coil radius and N is the number of turns of the coil winding. Thus the moving
coil can be modeled by the ideal energy-transfer transformer shown in Fig. 10.2b where the transforming
ratio
v
n = = 2π N B i (10.4)

In the model shown in Fig. 10.2b we neglect the electrical resistance and self-inductance of the voice-coil,
as well as any capacitance. We also neglect the mechanical mass and friction, as well as any restraint
from the lead-in wires. These effects can be respected by augmenting the ideal transformer with physical
elements representing these effects.

Figure 10.2: Moving-coil transducer.

10.3 Single-phase alternator.


A model of a simple single-phase alternator is easy to construct based on the same principle. A geometric
configuration of such a device is shown in Fig. 10.3a. Wires wrapped around a rotor or armature rotating
about the z axis are carried through a gap of substantially constant x-directed flux density B. The
component of the conductor velocity perpendicular to the flux is

ẏ = r ϕ̇ cos ϕ (10.5)

where r is the radius of the armature and ϕ̇ is its angular velocity. It is this component which must be
used in (10.2) for the voltage equation
v = B ℓ r ϕ̇ cos ϕ (10.6)
where ℓ is the effective length of all conductors cutting the flux lines with an average density value B.
Wires may be wound many times around the armature, and the total voltage generated is the sum of
all the voltages generated by each wrap. Note that voltage induced on one side of the rotor adds to that
on the other because although the direction of flux cutting is opposite on two ends of the rotor diameter,
the direction of current flow is also opposite on the two ends. The electric circuit can be completed with
clip rings as shown.
If current starts to flow through the rotor winding, a force given by (10.1) will act on it giving rise
to a torque opposing the rotation. The magnitude of the external torque required to hold the rotor in
equilibrium will be
τ = F r cos ϕ = −B ℓ r i cos ϕ (10.7)

This together indicates, that the alternator can be modeled by a pure transformer with the variable
ratio
ϕ̇ 1
n= = (10.8)
v B ℓ sin ϕ
as shown in Fig. 10.3b. Pole Cϕ represents the rotational motion of the rotor.
Real alternators have more complicated configurations than the one shown schematically in Fig. 10.3a
an may have sets of windings to generate three-phase power, but the basic modeling idea based on and
energy-transfer transformer still applies.
10.4. PERMANENT-MAGNET DC MACHINE 10-3

Figure 10.3: Single-phase ac rotational transducer.

10.4 Permanent-magnet dc machine


This machine consist of several windings situated in a magnetic field of constant flux density B much like
alternator windings but with a commutator (mechanical with carbon brushes, or electronic brushless)
that switches the windings into and out of the external electrical circuit as the armature rotates. As
shown schematically in Fig. 10.4a, a coil is connected for only a small fraction of revolution. During the
time any coil is connected, the cos ϕ terms in (10.6) and (10.7) are nearly unity. The equations therefore
are approximately
v = B ℓ r ϕ̇ and τ = −B ℓ r i (10.9)

which can be represented, as in Fig. 10.4b, by a simple ideal-energy transfer transformer with the trans-
forming ratio
ϕ̇ 1
n= = (10.10)
v Bℓr
There is some ripple in the characteristics for v versus ϕ̇ and i versus ϕ due to the finite number of coils
in a dc machine, but (10.10) are accurate enough for many design purposes.

Figure 10.4: Pure model of dc permanent magnet machine.

Equations (10.6) are usually written as

v = KE ϕ̇ and τ = −KT i (10.11)

where KE is the back-emf constant of the dc machine and KT is its torque constant . The designation of
the former constant comes from the custom to call the voltage v induced in the dc machine rotor back
electromotive force (back emf). It is important to note that both constants are exactly the same.
 Their
values are only equal when self-consistent unit system is used. Expressed in the SI units, KE V.s.rad−1
 
and KT N.m.A−1 . Thus both the constants can be replaced in (10.7) by a constant KEM called
electromechanical constant . This constant is determined by the machine dimensions and by the state of
its magnetisation.
Fig. 10.5a shows a configuration of a permanent magnet dc machine in which Ra and La represents
the armature circuit resistance and inductance, respectively. Naturally, we can add these elements also
to the basic ideal-transformer model of the machine together with the elements modeling mechanical
friction br of the rotor bearing and the rotor inertia J as shown in Fig. 10.4b.
10-4 MODULE 10. ELECTROMECHANICAL SYSTEMS

Figure 10.5: Dynamic model of a dc permanent magnet machine.

10.5 Movable-plate condenser


Pure energy-transfer transducers in the form of energy-transfer transformers and gyrators are not sufficient
for modeling some real electro-mechanical transducers which not only transfer, but also store energy when
either mechanical or electrical work is done on them. The stored energy can be recovered at a later time.
Furthermore, the energy removed can be taken out as either mechanical or electrical work without regard
to the form in which the energy was originally stored.
The first example of an energy-storing transducer is the movable-plate condenser shown in Fig. 10.6a.
The charge on the condenser is q, and the voltage across the plates is v. The displacement of the movable
plate is x, and the external force required to hold the movable plate in equilibrium (against the electrical
attraction of the fixed plate) is F .

Figure 10.6: (a) Movable-plate condenser, (b) pure electro-mechanical model.

We shall replace the condenser by a pure capacitor assuming that the phenomena like electrical
inductance and resistance or mechanical mass and friction could be modeled by additional physical
elements.
The constitutive relations for the movable-plate capacitor can be expressed as

v = v(q, x) F = F (q, x) (10.12)

In electrically linear case, the first of the relations (10.12) has the form v = q/C(x), where C(x) is the
capacitance when the movable plate is in the position x. In general, C(x) is a nonlinear function of x.
An important fact is that when q = 0, the force F must be zero for any x. If there is no charge, there is
no electric field, and hence no attractive force between the plates.
The electrical and mechanical power delivered to the capacitor in time dt represents the increase dWe
in its stored electrical energy, i.e.
∂We ∂We
dWe = vi dt + F ẋ dt = v dq + F dx = dq + dx (10.13)
∂q ∂x
Because of its conservative nature, the total electrical energy of the capacitor We (x, q) is obtained by
integrating (10.13) along any path from the datum configuration with x = 0 and q = 0. If we integrate
(10.13) from (x = 0, q = 0) to a configuration (q, x) along the rectangular path following first the x-axis
up to the point (q = 0, x), and then the q-axis up to (q, x), the electrical energy is, simply,
Z q
q dq q2
We (q, x) = = (10.14)
0 C(x) 2C(x)
10.6. MOVABLE-CORE SOLENOID 10-5

as there is no contribution from the segment along the x axis (where the force F is zero for any x).
Once the energy function We (q, x) is known, the constitutive relations (10.12) can be recovered by its
differentiation:
∂We q ∂We q2 C ′
v= = F = =− (10.15)
∂q C ∂x 2C 2
where the derivative C ′ (x) = dC(x)/dx.

Figure 10.7: (a) Energy-storing transformer, (b) its source implementation. (c) Energy-storing gyrator,
(b) its source implementation.

To compose the corresponding model let us rewrite (10.15) as

dq 1
i= = C(x) v̇ + K(x, v) ẋ F = − K(x, v) v (10.16)
dt 2
where K(x, v) = C ′ (x) v. Than it should be seen that a variable pure capacitor combined with a pure
energy-storing gyrator can model the electro-mechanical coupling as shown in Fig. 10.6b. Fig. 10.7d
shows that the energy-storing gyrator consistscan be formed from a controlled source of current, and of
a controlled source of force.

10.6 Movable-core solenoid


Fig. 10.8a shows a solenoid, or movable-core coil – the magnetic counterpart of the movable-plate con-
denser.

Figure 10.8: (a) Movable-core solenoid, (b) pure electro-mechanical model.

Replacing the coil by a pure inductor, we can express its constitutive relations in the form of equations
for its current i and the external force F (required to hold the core in equilibrium against the magnetic
attraction) in terms of the inductor flux linkage λ, and the movable core displacement x as

i = i(λ, x) F = F (λ, x) (10.17)


10-6 MODULE 10. ELECTROMECHANICAL SYSTEMS

If the relation between current and flux linkage is linear, the first equation becomes i = λ/L(x) where
L(x) is the inductance of the coil when the core is in the position x. When λ = 0, there is no magnetic
field, and hence no magnetic attraction. The force F is therefore zero for any x when λ = 0.
Using a procedure analogous to that used in the previous example, the total magnetic energy of
the inductor Wm (λ, x) can be obtained by integrating the increase dWm in its stored magnetic energy
corresponding to the electrical and mechanical power delivered to the inductor in time dt along a path
from the datum configuration (λ = 0, x = 0). In this way we find

λ
λ dλ λ2
Z
Wm (λ, x) = = (10.18)
0 L(x) 2L(x)

Again, the constitutive relations (10.17) can be recovered by differenting the energy function Wm (λ, x):

∂Wm λ ∂Wm λ2 L′
i= = F = =− 2 (10.19)
∂λ L ∂x 2L

where the derivative L′ (x) = dL(x)/dx.


Taking into consideration (10.19), we can assume that

di 1
v = λ̇ = L(x) + K(x, i) ẋ F = − K(x, i) i (10.20)
dt 2

where
K(x, i) = L′ (x) i (10.21)

Hence, the corresponding pure model can be of the form shown in Fig. 10.8b. It consists of a variable pure
inductor combined with an energy-storing transformer the implementation of which in terms of controlled
sources is shown in Fig. 10.7b.

10.7 Electromagnetic relay


This result is readily applicable to the electromagnetic relay shown in Fig. 10.9a. The relay model given
in Fig. 10.9b neglects saturation of the magnetic circuit, which is assumed to be infinitely permeable, and
ignores leakage and fringing fluxes. As the relay inductance

N 2 µ0 A N 2 µ0 A
L(x) = , the factor K(x, i) = − i
x x2

where N is the number of winding turns, x is the air-gap width, A is the cross-sectional area of the
air-gap, and µ0 is the air permeability. The electrical resistance of the relay winding, the mass of the
movable part, its friction, and the spring effect are represented in the model by physical elements.
Fig. 10.9c shows a relay model in which the magnetic circuit is represented in a more authentic way.
Coupling between the electrical and magnetic circuits is modeled there by a gyrator of gyration ratio
equal to N . The elements Pf and Pa represent permeances of the ferromagnetic part of the magnetic
circuit and of the air-gap, respectively. The magneto-mechanical coupling is modeled by an energy-storing
gyrator. The electrical voltage is replaced there by magnetic voltage, the electrical charge by magnetic
flux, and the electrical capacitance by permeance. In this case,

wa
K(x, wa ) = P ′ (x) wa =
µ0 A

where wa is the magnetic voltage across the air-gap.


The element Rf corresponds to the magnetic circuit losses. Saturation or hysteresis of the ferromag-
netic material can be taken into consideration within the element Pf , and also such effects like leakage
and fringing could be respected in this model in a straightforward manner by adding the relevant elements
to it.
10.8. SYNCHRONOUS RELUCTANCE MACHINE 10-7

Figure 10.9: (a) Electromagnetic relay, (b) electro-mechanical and (c) electro-magneto-mechanical mul-
tipole model.

10.8 Synchronous reluctance machine


The proposed ideal energy-storing transducer suits well also to rotational systems. Let us assume, that
the rotor and the stator of an elementary singly excited reluctance machine shown in Fig. 10.10a are
shaped so that the variation of the exciting winding inductance L is sinusoidal with respect to the rotor
position θ as shown in Fig. 10.10b, that is,
1
L(θ) = [Ld + Lq + (Ld − Lq ) cos 2θ ] (10.22)
2
where Ld and Lq is the direct-axis and quadrature-axis inductance, i.e. the maximum and minimum
value of the inductance L(θ), respectively.
In accordance with (10.21), the transducer factor K in the reluctance machine model given in
Fig. 10.10c can be computed as
dL
K(θ, i) = i(t) = i(t) (Lq − Ld ) sin 2θ

For an excitation
i = Im sin ωt
the model exhibits the synchronization phenomenon, i.e. a nonzero average torque is developed only at
the synchronous speed θ̇ = ωm such that at any instant θ = ωm t − δ, where δ is the rotor position at
t = 0, when the current i is also zero.
The model is augmented with additional physical elements representing the winding resistance, the
rotor inertia, and the bearing friction.

10.9 Variable-reluctance stepping motor


As an extension of the reluctance machine model, Fig. 10.11 gives a multipole model of a variable-
reluctance multiple-stack three-phase stepping motor which has a wound stator and an unexcited rotor
with three separate sets of teeth sections. The three sets of rotor teeth are magnetically independent.
10-8 MODULE 10. ELECTROMECHANICAL SYSTEMS

Figure 10.10: Electro-mechanical multipole model of a reluctance machine.

In general, for an n-phase motor, the stator teeth are displaced by 1/n of a tooth pitch from section to
section.
To form a multipole model of the stepping motor, we need again to know the form of its inductance.
In practice, the motor inductance L as a function of displacement θ may be again approximated by a
cosine wave, that is,
L(θ) = L1 + L2 cos rθ (10.23)
where L1 and L2 are constants and r is the number of teeth on each rotor section.
Now, in the case of a three-phase motor with an 10-teeth rotor, r = 10, let the equilibrium position
be the situation when phase A is energized. Then the inductance for phase A is given by

LA (θ) = L1 + L2 cos 10θ (10.24)

Substituting (10.24) into (10.21), we get

KA (θ, iA ) = −iA (t) L2 sin 10θ

Figure 10.11: Multipole model of a multiple-stack three-phase variable-reluctance stepping motor.


10.9. VARIABLE-RELUCTANCE STEPPING MOTOR 10-9

Phase B has its equilibrium position 120◦ behind the reference point. Therefore, the inductances and
the K-factors for phases B and C are then written as follows:

LB (θ) = L1 + L2 cos(10θ − 120◦)


LC (θ) = L1 + L2 cos(10θ + 120◦)
KB (θ, iB ) = −iB (t) L2 sin(10θ − 120◦)
KC (θ, iC ) = −iC (t) L2 sin(10θ + 120◦ )

The electrical circuits of the three phases are isolated so that each phase can be modeled in a separate
way. The total torque developed on the rotor is the algebraic sum of torques of the three phases. Thus

τ = τA + τB + τC (10.25)
10-10 MODULE 10. ELECTROMECHANICAL SYSTEMS
Module 11

Unified approach to modeling

11.1 Dynamic behavior of systems


Dynamic behavior of a dynamic system is governed by the flow of energy and matter between the system
and its surroundings as well as between subsystems of the system. Within the subsystems, energy and
matter may be stored and later released, or changed from one form to another. Therefore, before starting
any investigation of a system behavior we should clearly separate the system from its surroundings. Then
we should decide in which way to decompose the system into its disjoint subsystems.
Table 11.1 gives examples of configurations of very simple systems from different physical domains.
(We could easily extend our scope also to magnetic or thermal systems, but there is no equivalent of
the inductor or spring in these domains.) Each of the systems consists from four different parts. The
electrical system is formed by an RLC circuit supplied from an electronic source of electrical current. In
the fluid system, a pump driven by an electrical motor forces a constant flow of water from a reservoir
into an open tank. Water is returned back to the reservoir through a long pipe with a restriction at its
end. The rectilinear translational mechanical system is powered by a weight suspended on a cable via a
pulley. The cable is pulling a body connected to a hydraulic damper by a helical spring. The pulley is
of negligible moment of inertia and bearing friction. The rotational mechanical system consists from a
flywheel linked to a ventilator propeller by a long torsionally flexible shaft. The flywheel is driven by a
combustion engine.
Evaluation of dynamic behavior of real systems in practice – either by measurements, or by simulation
– is very often simplified by the following approximating assumptions:

1. the energy interactions between subsystems take place just in a limited number of interaction sites
formed by adjacent energy entries into the subsystems

2. the energy flow through each energy entry can be expressed by a product of two complementary
quantities called power variables

3. the total energy flow into a subsystem equals to the sum of the energy flows through all its energy
entries

The energy interactions in our examples are assumed to take place at the sites denoted by M, N
and and 0. In the case of the electrical system, these sites are made by interconnections of the energy
entries formed by electrical terminals of the system parts. In the fluid system, the energy entries are
formed by pipe inlets of the parts. The spring and the damper in the translational system interact via
common cross-sections of their links. We can assume that the cable, the spring and the body interact via
a cross-section of the body. We could easily identify similar interaction sites and energy entries of parts
in the rotational system

11.1.1 Power variables


Table 11.2 shows pairs of physical quantities that can be chosen as power variables in the simple systems
under discussion. The quantities in each power-variable pair complement each other in such a way that
one of them is a through variable whereas the other one is an across variable. The table gives also the

11-1
11-2 MODULE 11. UNIFIED APPROACH TO MODELING

Table 11.1: Examples of simple dynamic systems and their multipole models.
Configurations of real systems Multipole diagrams

M part3 N
M 3 L
1 part1 part2 part4
2 C G
4 N J
0

Part Real parts Physical elements


1 electrical energy supply source of electrical current
2 electrical condenser electrical capacitor
3 coil of wire electrical inductor
4 electrical resistor electrical conductor

M N
2
3 N part3
part1 L part4
1
M J part2 G
0 4 C

Part Real parts Physical elements


1 water pump driven by motor source of volume flow-rate
2 open water tank fluid capacitor
3 long pipe fluid inductor
4 water-flow restriction fluid conductor

M N

part3 part4
part1
L G
J part2
C

Part Real parts Physical elements


1 weight suspended on cable via pully source of force
2 mass of moving body inertor
3 helical spring pure spring
4 oil-filled damper pure damper
M N

part3
M part1 part4
1 N part2 L
3 J G
0 j 4 C
2
Part Real parts Physical elements
1 combustion engine source of torque
2 flywheel rotational inertor
3 long flexible shaft torsional spring
4 ventilator propeller rotational damper
11.1. DYNAMIC BEHAVIOR OF SYSTEMS 11-3

Table 11.2: Pairs of through- and across-variables


Power variables Energy variables
Physical through across through across
R R
domain i v idt vdt
Electrical electrical electrical flux
electrical current voltage charge linkage
[A] [V ] [C] [V.s]
magnetic magnetic magnetic
magnetic flow rate voltage flow
[W b/s] [A] [W b]
entropy
temperature entropy
thermal flow
[K] [J/K]
[W/K]
volume pressure
fluid or pressure volume
flow momentum
N/m2
   3
acoustic  3  m
N.s/m2
 
m /s
force velocity momentum displacement
translatory
[N ] [m/s] [N.s] [m]
angular angular angular
torque
rotary velocity momentum displacement
[N.m]
[rad/s] [N.m.s] [rad]

related pairs of energy variables, i.e. the through and across variables integrated with respect to time.
The vacancies in Table 11.2 correspond to quantities which have no physical meaning.
Note that the through and across variables differ by the way in which they are directly measured.
Through variables are measured between interacting adjacent energy entries by instruments included
between the entries after the entries were disconnected. Across variables are measured between dis-
tant entries by instruments applied to the entries without any disconnections (across variables between
adjacent entries are zero). This is illustrated by Fig. 11.6 showing a two-entry part connected to a system.
An important role is played by across variables measured with respect to an absolute across-variable
reference. In each of the examples given in Table 11.1, such a reference is denoted by 0. The absolute
reference of voltages in electrical systems – electrical ground – is embodied in the example by the system
chassis. In the fluid system, the free atmospheric pressure was chosen as the pressure reference. In
both the mechanical systems, the system frame fixed to the Earth surface is considered as the absolute
reference for velocities.
The third approximating assumption mentioned above can be expressed for a subsystem with n entries
as
n
X
P (t) = ij (t) · vj (t) (11.1)
j=1

where P is the total power consumed in the subsystem, ij is the through variable of the j-th entry, and
vj is the across variable of this entry with respect to the absolute reference, j = 1, 2, . . . , n. We thus

SYSTEM SYSTEM SYSTEM

A B
iB
PART A PART B
A PART B'
vBA B vA vB
(a) (b) (c)

Figure 11.1: Measuring of (a) an across variable between distant energy entries, (b) a through variable
between adjacent entries, (c) across variables of entries with respect to a reference.
11-4 MODULE 11. UNIFIED APPROACH TO MODELING

assume that the power consumed (dissipated or accumulated) in a subsystem is positive whereas the
power retrieved from the subsystem is negative.

11.1.2 Orientation of variables


The value of a variable is either positive or negative depending on the orientation of the variable with
respect to its polarity or direction reference. To denote the assumed positive orientation of variables
engineers usually use arrows. If the actual orientation of a variable is opposite to the direction of the
arrow, the value of the variable is taken as negative. Let us agree that the variable orientation is governed
by the following conventions.

Arrow Orientation Conventions

• Orientation of across variables in a system is indicated by arrows with empty arrowheads. These
arrows are always directed from the point of the assumed smaller value to the point of the assumed
larger value of the related across variable with respect to its absolute reference.

• Orientation of through variables is indicated by arrows with full arrowheads. The full-head arrow
direction is in agreement with the assumed positive direction of the motion of the associated medium.

Thus the empty-head arrow points in the direction of the assumed increasing value of the across
variable. The full-head arrows point in the direction of the assumed positive motion of a flow of positively
charged particles in the electrical domain, magnetic flux in the magnetic domain, heat flow in the thermal
domain, and flow of mass particles in the fluid or acoustic domain. In the mechanical domain, a full-head
arrow indicates the direction of the motion the related force or torque tends to cause.

11.1.3 Modeling with physical elements


Multipoles model energy interactions between subsystems or between subsystems and the system sur-
roundings under the approximating assumptions mentioned above. Each energy entry into a subsystem
is represented in its multipole model by a pole associated with a pair of the power variables. In the mul-
tipole model of a subsystem, each of the multipole poles stands for an energy entry into the subsystem. It
is associated with the pair of power variables related to the entry. Multipoles are represented graphically
by symbols with poles denoted by pins, i.e. by short line segments sticking out of the symbol outlines.
The multipole model of a complete system is portrayed by the multipole diagram consisting of graphical
symbols of the individual multipole models. The sites of energy interaction between adjacent energy
entries in a real system are represented in the related multipole model by its nodes. Pins denoting the
entries interacting at a site are interconnected with the node representing the site by line segments called
ideal links.
The drawings of systems shown in the left column of Table 11.1 can be readily converted into multipole
diagrams shown in the right column of the table. Each of the two-entry parts is represented there
by a symbol of an idealized physical element. Each of such elements models just one specific energy
phenomenon like energy generation, accumulation, or dissipation. For simplicity, only the dominant
dynamic effect of each part is respected in the diagrams in Table 11.1. More realistic models could be
set up for each part by combining several such elements.
Nodes M, N and 0 in the diagrams shown in Table 11.1 are related to the sites of energy interaction in
the real systems denoted by the respective characters. Reference nodes representing the absolute across-
variable references 0 are denoted in the multipole diagrams by graphical symbols shown in Table 11.3.

The links can be viewed as interconnections capable of transferring energy in both directions without
any dissipation, accumulation or delay. Depending on the related physical domain, the links can be
viewed either as idealized electricity-conducting wires or as fluid-conducting pipes, idealized shafts or
other mechanical links, etc. This does not mean, however, that real interconnections should be always
modelled in such an idealized way; they can be considered as independent subsystems with any of their
non-ideal features respected. This is the case of the long pipe segment in the fluid-system example, and
of the long torsionally flexible shaft in the rotational-system example in Table 11.1.
11.2. DISSIPATIVE AND ACCUMULATIVE ELEMENTS 11-5

Table 11.3: Absolute across-variable references


Physical domain Reference across variable Symbol

electrical voltage of the electrical ground

magnetic magnetic voltage at a reference point

thermal zero-point on a temperature scale

fluid or acoustic free-atmosphere pressure (usually)

mechanical reference-frame (angular) velocity

11.2 Dissipative and accumulative elements


11.2.1 Symbols of dissipative and accumulative elements
Table 11.4 gives graphical symbols for physical elements modeling dissipation and accumulation of energy
in multipole diagrams. The entries for thermal and magnetic inductors are empty in the table as no
corresponding physical phenomena are known. The notion of a resistor is not common in the mechanical
domain. Besides the element symbols specific for different physical domains, Table 11.4 gives also generic
symbols applicable to elements in any of the domains.
The physical elements are very useful approximations of dynamic effects associated with real sub-
systems. The elements are called pure models as each of them represents just one specific dynamic
effect. Depending on their physical domain, the elements model energy effects associated with the one-
dimensional motion of electrical or magnetic charges, fluids (liquids or gases), particles or bodies. The
thermal-domain elements are concerned about reversible heat-transfer effects.
Pure capacitors model accumulation of energy by the virtue of an across variable in the non-mechanical
domains. Inertors do the same in the mechanical domain, they model accumulation of inertial energy
associated with either rectilinear or rotational motion of a mass. Pure inductors and springs model
accumulation of energy by the virtue of a through variable. The energy accumulated in these elements
can be retrieved back at any time. With the exception of the thermal-domain elements, pure conductors,
resistors and dampers model dissipation of energy, i.e. the conversion of energy into heat. This energy
cannot be retrieved and returned back into the systems. Pure thermal conductors and resistors model
the increase of system entropy due to the flow of heat.

11.2.2 Constitutive relations of physical elements


As shown in Fig. 11.2, each of the physical elements is associated with a pair of power variables ie and
ve . Whereas ie is the element-through variable ‘flowing’ through the element, ve is the element-across
variable between the element poles. These two variables complement each other in the sense that their
product
Pe (t) = ie (t) · ve (t) (11.2)
expresses the power consumed (dissipated or accumulated) in the element. Thus, if the power is retrieved
from an element, its value is negative.

Figure 11.2: Variables associated with a generic physical element.

Each of the physical elements is characterized by a specific constitutive relation between the el-
ement variables. Constitutive relations for the dissipative and accumulative elements are given in
11-6 MODULE 11. UNIFIED APPROACH TO MODELING

Table 11.4: Physical elements dissipating or accumulating energy.


Type G R C L
Conductor/Damper Resistor Capacitor/Inertor Inductor/Spring
Non-mechanical Conductor Resistor Capacitor Inductor

electrical

magnetic

thermal

fluid or
acoustic

Mechanical Damper Inertor Spring

translational

rotational

generic

Table 11.5. The letter p denotes the element parameter p in these relations. Table 11.5 gives also
physical dimensions of the parameter p in different physical domains and their units. The fact that the
constitutive relation of the dissipative and accumulative elements are in the form of algebraic of ordinary
differential equations indicates that the elements belong to lumped-parameter models.
In the case of ideal elements, the parameter p is constant. The elements given in Table 11.4 need
not be ideal, however. DYNAST allows also for nonlinear and time-variable elements as well as for
elements the parameters of which are are controlled by variables or parameters of some other elements,
blocks or equations.
In general, the parameter p of a controlled nonlinear and time-variable element can be given by an
expression of the form
p = f (z1 , z2 , . . . , ż1 , ż2 , . . . , t) (11.3)
where z1 , z2 , . . . are elementcontrolling variables or parameters, ż1 , ż2 , . . . are their time derivatives, and
t denotes time.
Despite the fact that all the physical elements are two-pole models, one-pin symbols are given in
Table 11.4 for thermal, fluid or acoustic capacitors as well as for the translational and rotational inertors.
According to physical laws governing accumulation of energy by these elements, their element-across
variable ve must be always considered with respect to the related reference. In other words, one pole of
11.3. SOURCES OF ENERGY 11-7

Table 11.5: Parameters of dissipative and accumulative physical elements.

Element Conductor or damper Resistor Capacitor or inertor Inductor or spring

d d
Constitutive relation ie = p ve ve = p ie ie = p dt ve ve = p dt ie

conductance resistance capacitance inductance


electrical [S] [Ω] [F ] [H]
conductance resistance permeance
magnetic [H]
[Ω] [S]
entropic entropic entropic
thermal conductance resistance capacitance)
W/K 2
   2 
K /W [J/K]

fluid or acoustic conductance resistance capacitance  inertance


P a.s/m3
 3
m3 /(P a.s) P a.s2 /m3
   
m /P a
mechanical damping mass compliance
translatory [N.s/m] [kg] [m/N ]
torsional moment of torsional
mechanical
rotary
damping  intertia compliance
m2 kg/rad

[N.m.s/rad] [rad/(N.m)]

these elements must be always coalesced with a reference node. Only one pin needs to be shown in the
graphical symbols of these elements to simplify the multipole diagrams.

11.3 Sources of energy


11.3.1 Independent sources of energy
Besides the energy storing and dissipating elements we need still another type of physical elements to
be able to set up multipole models of dynamic systems – pure sources of energy. Two limiting cases
are considered in which either an across variable or a through variable is specified independently of the
amount of energy drawn from or delivered into the source – a source of across variable and a source of
through variable. Our symbols for such sources are given in Table 11.6. We shall use the generic sources
in all non-mechanical domains. The last row in Table 11.6 shows that the parameter p in the constitutive
relations of sources equals directly to the source nominal variable, either ie or ve .
In general, pure sources are either independent or controlled by various variables or parameters. The
independent sources can be ideal or time-variable. An ideal independent across-variable source
ve (t) is defined by the constituency relation
ve (t) = V (11.4)
and, similarly, an ideal independent source of through variable ie (t) is described as
ie (t) = I (11.5)
where V and I are known constant parameters.
The independent sources are usually used to model external energy reservoirs large enough relative to
the amount of energy they supply to the system or absorb from it without undergoing any change either
in across or through variable. Fig. 11.3a shows a system and its surroundings modelled by twopoles. If
the surroundings is capable of delivering energy into the system with negligible variations of the through
variable i(t) involved in the interaction, the surroundings can be modelled by an across-variable source as
shown in Fig. 11.3b. If, on the other hand, the surroundings delivers energy without noticeable fluctuations
of the across variable v(t), the model configuration with a through-variable source of Fig. 11.3c can be
utilized.
Examples of real sources that approximate independent sources:
• A low-resistance electrical battery supplying a low-consumption electronic device is very often
considered as an independent source of constant voltage. Energy supply in the form of electrical
mains is usually modelled as an ideal source of sinusoidal voltage.
11-8 MODULE 11. UNIFIED APPROACH TO MODELING

Table 11.6: Independent sources of energy.


Type J E
Generic through-variable source across-variable source

Mechanical
source of force source of velocity

translational

source of torque source of angular velocity

rotational

Constitutive relation p = ie p = ve

Figure 11.3: (a) Interaction between a system and its surroundings modelled by (b) across-variable source,
or (c) through-variable source.

• An independent source of magnetic voltage can be used to model generation of magnetic field in a
permanent magnet.

• In the analysis of the heat dissipation from a system in a room, the air in the room can be treated
as an independent source of temperature if the heat transfer from the system to the room air is not
large enough to have a noticeable effect on the room air temperature.

• Most of the hydraulic pumps are usually modelled as independent sources of volume flow.

• A large accumulator supplying fluid to a small fluid-power system may be replaced by an indepen-
dent source of pressure.

• A sufficiently powerful motor driving a light mechanism represents from the point of view of the
mechanism an independent source of angular velocity.

• When modeling motion of a body close to the earth surface, the effect of the earth gravitational
field on the body can be respected by an independent source of gravitational force Fg = g · m, where
g is the gravitational constant and m is the mass of the body.

It is assumed in each of the examples given in Table. 11.1 that the system under our investigation
consists of parts 2, 3 and 4 while the part 1 is considered there as the system surroundings. For this
reason, each of the parts 1 is modelled by an independent source of through variable. If modeling of a
real source or the system surroundings by an independent source alone is inadequate, the independent
source is combined with other physical elements.
11.3. SOURCES OF ENERGY 11-9

11.3.2 Ideal connections, indicators and sensors


Also zero-valued independent sources deserve our attention. An independent source of zero across variable
is equivalent to an ideal closed connection, or simply, an ideal connection. This is a two-pole element
the across variable ve of which is identically zero irrespective of its through variable ie , i.e., its constituency
relation is ve = 0. Similarly, a two-pole element makes an ideal open connection if its through variable
i is identically zero irrespective of its across variable v, i.e., its description is i = 0. Obviously, such
an element is equivalent to an independent source of zero through variable. We shall find not only the
ideal closed connections, but also the open connections, very useful when modeling a changing system
structure. Fig.11.4 shows characteristics of these elements the v-i and i-v planes.

Figure 11.4: Characteristic of ideal (a) closed and (b) open connection.

Note also, that the zero-valued independent sources can be used to imitate measuring instruments
and sensors of power variables. An ideal through-variable indicator models ideal through-variable
measuring or sensing devices. It behaves like an ideal connection the through-variable of which corre-
sponds directly to the indicated quantity. Similarly, an ideal across-variable indicator acts as an ideal
open connection and indicates its own across variable. Therefore, the ideal through-variable indicator is
equivalent to an independent source of zero across variable, whereas the ideal across-variable indicator is
equivalent to an independent source of zero through variable.

Figure 11.5: Graphical symbols for an ideal (a) across-variable and (b) through-variable indicator.

In cases when we shall need to stress the presence of ideal measuring instruments in a system model,
we shall use the graphical symbols shown in Fig. 11.5. Each of the symbols for the ideal indicators
represents a two-pole with its +pole at the top.
Fig. 11.6 shows configurations in which ideal indicators are used to ‘pick-up’ power variables of a
generic physical element in a system. The instrument orientation with respect to the element orientation
was chosen in such a way that the variable polarities are in mutual agreement.

Figure 11.6: Configurations for indicating (a) element-across variable and (b) element-through variable.
11-10 MODULE 11. UNIFIED APPROACH TO MODELING

11.3.3 Controlled sources


Many physical devices controlling a flow of energy or transforming one form of energy into another form
may often be adequately modelled by means of controlled sources. The element-across variable ve or
element-through variable ie of these two-pole physical elements is a function of some other variables in
the system model. In ideal controlled sources, this function is proportionality. Table 11.7 lists four types
of ideal controlled sources, controlled either by an across variable vc or by a through variable ic . No
energy input is required to control the sources. The symbol given in Table 11.6 will be used throughout
this text both for independent and controlled sources.

Table 11.7: Ideal controlled sources.


Source of Controlled by Relation Parameter p
across variable across variable ve = p · vc across-variable transfer ratio
across variable through variable ve = p · ic transresistance
through variable through variable ie = p · ic through-variable transfer ratio
through variable across variable ie = p · vc transconductance

In general, the controlled sources of through variables are characterized by the constitutive relation

ie = f (z1 , z2 , . . . , ż1 , ż2 , . . . , t)

The counterparts, controlled sources of across variables are characterized by the relation

ve = f (z1 , z2 , . . . , ż1 , ż2 , . . . , t)

11.4 Orientation of physical elements


11.4.1 Element Polarity Convention
To simplify the polarity determination of variables associated with physical elements as much as possible,
we can assume that the variable orientation is fixed to each element symbol once for ever. For the
element-polarity reference, we are associating one of the poles of each element with the + sign, and the
other pole with the − sign. The − sign is not shown in the element symbols given in Table 11.4, however,
in order to simplify the notation. In the case of the elements with an asymmetric symbol even the + sign
is omitted for the same reason. We will always assumed that the pin of asymmetric symbols shown in
Table 11.4 in the upper position is associated with the + sign.
To interrelate orientation of physical elements and their variables, let us adopt the following conven-
tions:

• The element-across variable ve is the across variable of the element +pole with respect to its −pole
(or to the absolute reference if the second pin is not shown in the element symbol).

• The positive orientation of the element-through variable ie is always directed from the +pole towards
the −pole.

Fig. 11.2 shows that according to our conventions the empty-head arrow for ve is always directed
from the −pole towards the +pole of the element, whereas the full-head arrow for ie is always directed in
the opposite way. The empty-head arrow for the pole variable v+ and v− is always directed away from
the absolute reference towards of the +pole and −pole, respectively. Note, however, that once we have
interrelated the variable orientation with the element polarity, we do not need to use the arrows at all.

11.4.2 Orientation of non-mechanical elements


Fig. 11.7a shows a real subsystem with two energy entries of non-mechanical nature, Fig. 11.7b gives a
generic symbol for a physical element modeling such a subsystem. The orientation of the element was
chosen in such a way that its +pole represents the entry A and its −pole the entry B of the subsystem.
Then, in conformity with the Element Polarity Convention, the element-across variable ve corresponds
11.4. ORIENTATION OF PHYSICAL ELEMENTS 11-11

to the across variable vBA of the entry B with respect to the entry A as indicated in Fig. 11.7a. Similarly,
the orientation of the element polarity with respect to the entries A and B determines orientation of the
element-through variable ie . It corresponds in this case to the through variable entering the subsystem
at the entry A and leaving it at the entry B.
If we had chosen the opposite orientation of the element with respect to the subsystem entries, also
the orientation of the variables ie and ve would be opposite. Note that the presumption (11.2) is satisfied
regardless of the chosen element orientation with respect to the entries.

Figure 11.7: (a) Subsystem with two energy entries of non-mechanical nature, (b) its generic-element
model.

11.4.3 Orientation of mechanical elements


Fig. 11.8 shows a two-entry subsystem in a translational motion along the axis x and its model in the form
of a generic physical element. The energy entries A and B of the subsystem Fig. 11.8a are represented
by the element poles Ax and Bx of the element in Fig. 11.8b. Once the direction of the x-axis has
been chosen from A towards B, it is advantageous to choose the orientation of the element as shown in
Fig. 11.8b to simplify the variable orientation. Then, in conformity with the Element Polarity Convention,
the element-across variable ve represents the velocity ẋBA of the entry A with respect to the entry B.
Note that in this case ve is positive if the subsystem is elongating and negative if it is shortening.
To satisfy the presumption (11.2), we must assume that the element-through variable ie represents
the subsystem internal force counterbalancing an external force stretching the subsystem. Thus, a pos-
itive ie indicates that the subsystem is in tension, whereas a negative ie indicates a compression of the
subsystem. Had we chosen the opposite orientation of the element with respect to the direction of the
related coordinate, also the orientation of the two variables ve and ie would change.

Figure 11.8: (a) Subsystem with two energy entries in translational motion, (b) its generic-element model.

Fig. 11.9 shows the similar situation for a subsystem with entries in rotational motion. In summary,
the mutual orientation of variables associated with a mechanical subsystem and its elemental model
must be related to the mutual orientation of the subsystem coordinate axes and the element polarity.
Whenever possible, we shall always prefer the mutual subsystem-element orientation shown in Fig. 11.8b
and Fig. 11.9b.

11.4.4 Orientation of sources of energy


The Element-Polarity Convention applies also to sources. The source of force in Fig. 11.10b models the
translational actuator shown in Fig. 11.10a providing a pair of forces acting between the attachment
points A and B. The forces F and −F making up the pair have a common line of action but are in
opposite directions. If F is positive, the forces pull on the points A and B, and if F is negative the forces
push on the points. Fig.11.10b and c show sources of force of different polarities with respect to the
mutual orientation of the modelled actuator and the x axis. Note that the sign of the force is related to
the source polarity.
11-12 MODULE 11. UNIFIED APPROACH TO MODELING

(a) (b)
Figure 11.9: (a) Subsystem with two entries in rotational motion, (b) its generic physical-element model.

Figure 11.10: (a) Force actuator, (b) and (c) corresponding sources of force.

Source of velocity imposes a relative motion of a specified velocity between two points it represents.
Fig. 11.11a shows point A that is in a rectilinear motion along the x-axis of a given absolute velocity
ẋA (t). In Fig. 11.11b the x motion of the point A is represented by node Ax , the actuator giving the
velocity ẋA (t) to the point A an ideal source of velocity modeling motion of a point A along the x-axis
with known absolute velocity ẋA (t).
The source of velocity shown in Fig. 11.11b models imposing a relative motion of point B with respect
to point A with known relative velocity ẋBA (t) along the x-axis as shown in Fig. 11.11a.

(a) (b)
Figure 11.11: (a) Moving points, (b) source of velocity modeling an imposed motion of the points.

11.5 Configurations of physical elements


11.5.1 Setting up a physical-element model
• Sketch the geometric configuration of your system

• Define the system and its surroundings and denote by identifiers the sites of energy interaction
between them

• Decompose the system into subsystems and denote the sites of energy interaction between the
subsystems

• Associate all the sites of mechanical-energy interaction with a suitable coordinate system

• Start sketching the multipole model of your system by placement of nodes representing the indi-
vidual sites of energy interaction

• Add reference nodes using the symbols given in Table 11.3

• Choose a suitable physical-element model for each subsystem and place it between the nodes cor-
responding to the subsystem interaction sites
11.5. CONFIGURATIONS OF PHYSICAL ELEMENTS 11-13

• Orient the polarity of each mechanical element in conformity with the chosen coordinate system

• Repeat the previous two steps for the surroundings

• You may add ideal indicators for the across and through variables you are interested in

Though the symbols were chosen to remind real subsystems you should never forget about the dif-
ference between a real subsystem and its model. As a matter of fact, no real subsystems behave exactly
like the physical elements. The elements are, however, very useful first-order approximations of dy-
namic effects taking place in real subsystems. If modeling of a subsystem by a single element appears to
be inadequate, a more realistic subsystem model can be constructed using several physical elements as
standardized building bricks.

11.5.2 Forbidden source configurations


The independent sources discussed so far can be used in a very flexible way, yet not without any restric-
tions. Fig. 11.12a shows a system model with three independent across-variable sources forming a close
loop. According to Postulate of Compatibility the element-across variables of the sources must satisfy
the equation
v1 (t) − v2 (t) − v3 (t) = 0 (11.6)

for any t. Otherwise the variable would be inconsistent. If, however, (11.6) is satisfied, any one of
the three across-variable sources can be omitted without any impact on the system-model dynamics.
Therefore, if (11.6) is not satisfied, the equations characterizing the system configuration in Fig. 11.12a
have no solution. If (11.6) is satisfied, the equations are singular unless the superfluous source is removed
from the system model.

Figure 11.12: Forbidden (a) loops of across-variable sources and (b) cut-sets of through-variable sources.

The three through-variable sources in Fig. 11.12b form a cut-set, or – in other words – if the three
sources are removed from the system model, the model gets separated into two isolated parts. At the
same time, according to Postulate of Continuity the element-through variables of the sources must for
any t satisfy the equation
i1 (t) − i2 (t) + i3 (t) = 0 (11.7)

Then again, if (11.7) is satisfied any one of the sources can be omitted in the configuration without
affecting the system-model behavior. Also in this case, unless the superfluous source is removed from the
system model, the equations characterizing the system are either having no solution or they are singular.
This observation, that can be easily proven in a general way, results in the following conclusion:
Loops of across-variable sources and cut-sets of through-variable sources are forbidden in the system-
model configurations.
Fig. 11.13 and Fig. 11.14 give examples of permitted and forbidden source configurations. Explain
the reason why the configurations in Fig. 11.14 are forbidden and try to submit them to DYNAST to see
what kind error messages DYNAST will give you.
11-14 MODULE 11. UNIFIED APPROACH TO MODELING

Figure 11.13: Examples of permitted configurations.

Figure 11.14: Examples of forbidden configurations.

11.6 Energy transducers


11.6.1 Energy-transfer transducers
In general, transducers are devices converting energy from one form to another. For example, motors
and other actuators convert electrical, fluid or some other energy into the mechanical one, sensors usually
convert different forms of energy into the electrical form, etc.
In this section we are going to introduce pure transducers which are multipoles modeling energy
conversion without any energy loss. They allow not only for very easy and unified way of modeling
various real transducers converting energy between different energy domains, but they are also convenient
for modeling various couplings within the individual energy domains. Combined with two-pole physical
elements, pure transducers make fundamental ‘bricks’ in our kit from which realistic models of very
diverse dynamic systems can be set up in an extremely easy way.
Pure transducers are subdivided into transducers which can store energy and those which only transfer
energy. An energy-transfer transducer is a multipole for which the instantaneous power out always equals
the instantaneous power in without dissipating or accumulating any energy. The pure energy-transfer
transducers are either pure energy-transfer transformers or pure energy-transfer gyrators.

Figure 11.15: Generic pure transducer.

Let us consider the four-pole shown in Fig. 11.15 which consists of two generic two-pole elements
which are mutually controlled by each other’s element variables. Assuming that the coupling is pure, i.e.
11.6. ENERGY TRANSDUCERS 11-15

that the total power lost or accumulated in the coupling is zero, the across and through variables va (t)
and ia (t) of the element a together with the analogous variables vb (t) and ib (t) of the element b must
justify the energy-transfer relation

ia (t) · va (t) + ib (t) · vb (t) = 0 (11.8)

With regards to signs of the variables, we have respected here the Element-Variable Polarity Convention.

11.6.2 Energy-transfer transformers


Let us assume, that the coupling between variables of the generic elements a and b in Fig. 11.15 can be
characterized by the describing relation

f qa (t), qb (t) = 0 (11.9)

where f (.) is a time-independent function and qa (t), qb (t) are time-integrated element-through variables
of the elements a and b, respectively (in mechanics such couplings are called holonomous scleronomous).
The variables come possibly but not necessarily from different energy domains.
If the function f (.) is differentiable, we can apply the chain rule to differentiate (11.9) with respect
to time to obtain
∂f dqa ∂f dqb
· + · =0 (11.10)
∂qa dt ∂qb dt
Taking into consideration that dqa /dt = ia (t) and dqb /dt = ib (t), (11.10) can be rewritten as

na · ia (t) + nb · ib (t) = 0 (11.11)

where na = ∂f /∂qa and nb = ∂f /∂qb .


By combining (11.11) with (11.8) we are arriving at the constitutive relationship of the pure energy-
transfer transformer which is usually written in the explicit matrix form
     
va (t) 0 n ia (t)
= · (11.12)
ib (t) −n 0 vb (t)

where
na
n= [−] (11.13)
nb
is the dimensionless coupling ratio of the energy-transfer transformer.
To arrive at the same pure energy converting model, Instead from (11.9) we could have started the
derivation of the energy-transfer transformer constitutive relation from the describing relation

g λa (t), λb (t) = 0 (11.14)

specifying a coupling between the time-integrated element-across variables λa (t), λb (t) of the elements a
and b. Following a similar procedure, in this case we would arrive at the constitutive realtion
     
ia (t) 0 m va (t)
= · (11.15)
vb (t) −m 0 ib (t)

where
nb 1
m=− =− (11.16)
na n
Obviously, m is a negative reciprocal of the coupling ratio n defined by (11.13).
Fig. 11.16 presents examples of pure energy-transfer transformer implementations in which the generic
elements a and b are replaced by controlled sources. The source combinations shown in Fig. 11.16a and
b correspond directly to (11.12) and (11.15). The source combinations given in Fig. 11.16c and d must
be considered simultaneously with (11.11) and 11.8). Thanks to the implicit form of the equations, the
latter transformer models are more flexible from the point of view of their interconnection with other
multipoles in a system models.
Graphical symbols which we shall use for pure energy-transfer transformers are indicated in Fig. 11.17.
In the case of the symbol with the slash sign shown in Fig. 11.17a the parameter k equals to the coupling
ratio (11.13), i.e., k = na /nb , whereas in the case of the symbol with the backslash sign shown in
11-16 MODULE 11. UNIFIED APPROACH TO MODELING

Figure 11.16: Implementation of energy-transfer transformers in terms of controlled sources.

Figure 11.17: Graphical symbol for energy-transfer transformers with (a) k = na /nb and (b) k = nb /na .

Fig. 11.17b the parameter k is defined by the reciprocal expression, i.e., k = nb /na . We are assuming,
that the index a is always associated with the left-hand pair of poles of the symbol, whereas the index
b always belongs to the right-hand pole-pair. For this reason, the symbol must be always situated
horizontally in a multipole diagram.
An ideal energy-transfer transformer is the pure transformer the coupling ratio of which is a con-
stant. We shall encounter numerous examples, however, in which the ratios are functions of some system
variables or parameters.

11.6.3 Load conversions with transformers

Figure 11.18: (a) Cascade interconnection of energy-transfer transformers, (b) one-transformer equivalent.

Fig. 11.18a shows a ‘cascade’ interconnection of energy-transfer transformers of coupling ratios n1 and
n2 . A simple analysis of the total transformation of across and through variables in the interconnection
indicates that the cascade can be replaced by only one transformer with the coupling ratio n1 n2 shown
in Fig. 11.18b.
Fig. 11.19 indicates the way in which series or parallel combinations of physical elements connected
to a pole-pair of an energy-transfer transformer are ‘seen through’ the transformer, or referred to at the
other transformer pole-pair. Z resp. Y is an impedance resp. admittance of the generic element, E resp.
J is an independent source of an across resp. through variable. This is the principle behind utilizing
various transducers and couplings to convert the amount of an actuator load at the price of changed
related across and through variables.
11.6. ENERGY TRANSDUCERS 11-17

Figure 11.19: Load ‘seen through’ an energy-transfer transformer.

11.6.4 Energy-transfer gyrators


To model electro-magnetic, fluid-mechanical or acoustic-mechanical transducers, for instance, we need
still another kind of a pure converter – the energy-transfer gyrator. The corresponding constitutive
relation can be derived in a similar way that we used for the energy-transfer transformer. Let us assume
this time, that the coupling between variables of the generic elements a and b in Fig. 11.15 can be
characterized by the describing relation

f λa (t), qb (t) = 0 (11.17)
where f (.) is a time-independent function and λa (t) is a time-integrated element-across variable of the
element a whereas qa (t) is a time-integrated element-through variable of the element b. The variables
come possibly but not necessarily from different energy domains.
If the function f (.) is differentiable, we can apply the chain rule to differentiate (11.17) with respect
to time to obtain
∂f λa ∂f qb
· + · =0 (11.18)
∂λa dt ∂qb dt
Taking into consideration that dλa /dt = va (t) and dqb /dt = ib (t), (11.18) can be rewritten as
na · va (t) + nb · ib (t) = 0 (11.19)
where na = ∂f /∂λa and nb = ∂f /∂qb .
By combining (11.19) with (11.8) we are arriving at the constitutive relationship of the pure energy-
transfer transformer      
ia (t) 0 S va (t)
= · (11.20)
ib (t) −S 0 vb (t)
where
na
S= (11.21)
nb
is the gyrating ratio of the energy-transfer gyrator. Note that S is a transconductance and its physical
dimension is energy-domain dependent.

Figure 11.20: Implementation of energy-transfer gyrators in terms of controlled sources.

Fig. 11.20 presents examples of pure energy-transfer transformer implementations in which the generic
elements a and b are replaced by controlled sources. The source combination shown in Fig. 11.20a
corresponds directly to (11.20), Fig. 11.20b corresponds to the inversion of (11.20).
11-18 MODULE 11. UNIFIED APPROACH TO MODELING

Figure 11.21: Graphical symbols for energy-transfer gyrators.

Graphical symbols which we shall use for pure energy-transfer transformers are indicated in Fig. 11.21.
An ideal energy-transfer gyrator is a pure gyrator the gyrating ratio of which is a constant. We shall
encounter numerous examples, however, in which the ratios are functions of some system variables or
parameters.

11.6.5 Load conversions with gyrators

Figure 11.22: (a) Cascade interconnection of an energy-transfer gyrator and transformer, (b) gyrator
equivalent.

Fig. 11.22a shows a ‘cascade’ interconnection of an energy-transfer gyrator of gyrating ratio S and
an energy-transfer transformer of coupling ratio n. A simple analysis of the total transfer of across and
through variables through the interconnection indicates that the cascade can be replaced by only one
gyrator with the gyrating ratio Sn shown in Fig. 11.22b.
Fig. 11.23 indicates the way in which series or parallel combinations of physical elements connected to
a pole-pair of an energy-transfer gyrator are ‘seen through’ the gyrator, or referred to at the other gyrator
pole-pair. Note that gyrators convert inductors or springs into capacitors or inertors and capacitors or
inertors into inductors or springs. This principle can be utilized both in system analysis as well as in
system construction.
11.6. ENERGY TRANSDUCERS 11-19

Figure 11.23: Loads ‘seen through’ an energy-transfer gyrator.


11-20 MODULE 11. UNIFIED APPROACH TO MODELING
Module 12

Formulation of system equations

12.1 Full set of equations


Values or waveforms of variables in a system multipole model can be obtained by solving a set of equations
characterizing the model. To formulate such equations, we must take into our consideration the following
three types of relations.
(a) Constitutive relations characterizing each of the individual multipoles in the model.
(b) Postulate of continuity reflecting the physical laws governing the balance of energy and mass at
the individual sites of interaction in the system. The laws are given for different physical domains in
Table ??. According to this postulate the sum of all the elemental through variables ik at a node
X
ik = 0 (12.1)
k

assuming that ik > 0 if ik leaves the node, and ik < 0 if ik enters the node.
(c) Postulate of compatibility resulting from the laws respecting the geometric connectedness of real
systems given in Table ??. According to this postulate the sum of all the elemental across variables vk
along a closed oriented loop in the diagram
X
vk = 0 (12.2)
k

where vk > 0 if the vk direction is in agreement with the loop orientation, and vk < 0 if the vk direction
is opposite to the orientation of the loop.
To demonstrate the principles of such equation formulation, let us start with the examples presented
in Table 11.1 already.
(a) Constitutive relations of physical elements involved in the examples are given in Table 12.1 in
three different forms where J, C, L and G are parameters of the elements, ik is the through variable, vk
is the across variable of the k-th element in each of the diagrams. Let us choose the implicit form:

Table 12.1: Constitutive relations of the elements in Table ??.


Part Physical elements Implicit relations Relations for i Relations for v
1 through-variable source i1 − J = 0 i1 = J —
i2 − C dv d
v2 = C1 i2 dt
R
2 capacitor or invertor dt = 0
2
i2 = C dt v2
v3 − L di 1 d
R
3 inductor or spring dt = 0
3
i3 = L v3 dt v3 = C dt i3
1
4 conductor or damper i4 − Gv4 = 0 i4 = G v4 v4 = G i4

i1 − J = 0
d
i2 − C dt v2 = 0
(12.3)
d
v3 − L dt i3 = 0
i4 − G · v4 = 0

12-1
12-2 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

(b) Fig. 12.1 shows the graph of the multipole diagrams in Table 11.1. The graph nodes are identical
with the nodes of the diagrams. Each of the diagram elements is represented in the graph just by
a line segment. In Fig. 12.1(a) the line segments are associated with arrows indicating orientation of
the elemental through variables (fixed to the orientation of the individual elements). According to the
postulate of continuity (12.1) for the elemental through variables at nodes M and N

node M: i1 + i2 − i3 = 0
(12.4)
node N: i3 + i4 = 0
(c) Fig. 12.1(b) shows two independent loops in the graph. The loop a is made up by elements 1 and
2, and the loop of elements 2, 3 and 4. Arrows indicate the chosen orientation of the loops with respect
to the orientation of the elemental across variables. According to the postulate of compatibility (12.2),
the elemental across variables along the loops a and b

loop a: v1 − v2 = 0
(12.5)
loop b: v2 + v3 − v4 = 0

i3 v3
M N M N

i2 i4 b
v2 v4
i1 v1 a

0 (a) 0 (b)

Figure 12.1: Graph of the multipole diagrams in Table 11.1 with arrows indicating orientation of elemental
variables and chosen loops.

The four constitutive relations (12.3) together with the two continuity relations (12.4) and two com-
patibility relations (12.5) form eight equations for eight unknown variables. In this case, the number of
equations can be reduced by one after substituting i1 = J to (12.4).
In general, for a model composed from m physical elements, the described procedure, called also the
tableau formulation, results in 2m differential equations to be solved to obtain 2m unknown system
variables.

12.2 Reduced equations


12.2.1 Nodal formulation
For practical reasons, engineers tend to decrease the number of equations to be solved simultaneously.
Nodal formulation is one of the procedures resulting in such reduced equations.
A systematic method of carrying out the nodal formulation for a model of n non-reference nodes (i.e.
nodes that were not chosen as reference nodes) is following:
Step 1. For each of the n non-reference nodes write the continuity equation (12.1).
Step 2. Express all the elemental through variables in the continuity equations in terms of the elemental
across variables using the constitutive relations of the individual elements.
Step 3. Write the elemental across variables in terms of the nodal across variables, i.e. the across
variables of the non-reference nodes relative to the reference node.
Step 4. Solve the resulting n equations for the n nodal across variables.
Step 5. Find the elemental across variables in the model as a difference of two of the nodal across
variables.
Step 6. Express the through variables of the elements using their constitutive relations.
12.2. REDUCED EQUATIONS 12-3

Examples

These steps can be illustrated again with the system models shown in Table 11.1.
Step 1. This step results in equations stated already in (12.4).
Step 2. Substituting the explicit constitutive relations for the elemental through variables from Table 12.1
into (12.4) gives

d
v2 − L1 v3 dt − i3 (0) = 0
R
J + C dt
(12.6)
1
R
L v3 dt + i3 (0) + Gv4 = 0

Step 3. The elemental across variables expressed in terms of the nodal across variables vM and vN are

v1 = v2 = vM
v3 = vN − vM (12.7)
v4 = vN

After substituting (12.7) into (12.6) we obtain

d
vM − L1 (vN − vM )dt
R
C dt = i3 (0) − J
(12.8)
1
R
L (vN − vM )dt + GvN = −i3 (0)

where the known variables were shifted to the right-hand sides of the equations.
Step 4. Using the nodal formulation, we have arrived at only two equations (12.8) to be solved simul-
taneously for the two unknown nodal across variables vM and vN .
Step 5. The elemental across variables are expressed in terms of the nodal across variables as in (12.7).
Step 6. The elemental through variables can be evaluated using the explicit constitutive relations given
in Table 12.1.
Note, however, that the nodal formulation cannot be applied directly to models with physical ele-
ments like sources of across variable, for example, as the explicit constitutive relation for their through
variables necessary in Step 2 of the formulation does not exist. Also, the integro-differential form of the
equations (12.8) does not suit to common computer computational routines for simultaneous solving of
the equations.

12.2.2 Loop formulation


An alternative formulation method for formulating equations well known from electrical-engineering text-
books is the loop formulation (or mesh formulation for planar models). The number of equations resulting
from this formulation is ℓ for a model with ℓ independent loops. A systematic method of carrying out
the loop formulation is the following.

Step 1. For each of the ℓ loops write the compatibility equation (12.2).

Step 2. Express all the elemental across variables in the compatibility equations in terms of the elemental
through variables using the constitutive relations of the individual elements.

Step 3. Write the elemental through variables in terms of the loop through variables.

Step 4. Solve the resulting ℓ equations for the ℓ loop through variables.

Step 5. Find the elemental through variables in the model as a difference of two of the loop through
variables.

Step 6. Express the across variables of the elements using their constitutive relations.
12-4 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

Example
To illustrate these steps let us apply them to the system models shown in Table 11.1 in which the source
of through variable was replaced by the source of across variable characterized by the constitutive relation

v1 = E (12.9)

Step 1. The the compatibility equations for this example are already in (12.5).
Step 2. Substituting the explicit constitutive relations for the elemental through variables from Table 12.1
into (12.4) gives

1
R
E− C i2 dt − v2 (0) = 0
(12.10)
1 d 1
R
C i2 dt + v2 (0) + L dt i3 + G i4 = 0

Step 3. The elemental through variables expressed in terms of the loop through variables ia and ib are

i1 = −ia
i2 = ia − ib
(12.11)
i3 = −ib
i4 = ib

After substituting (12.11) into (12.10) we obtain

− C1
R
(ia − ib )dt = −E + v2 (0)
(12.12)
1 d 1
R
C (ia − ib )dt − L dt ib + G ib = −v2 (0)
where the known variables were shifted to the right-hand sides of the equations.
Step 4. Using the loop formulation, we have arrived at only two equations (12.12) to be solved simul-
taneously for the two unknown nodal across variables ia and ib . (By adding the second equation to the
first one the number of equations to be solved is reduced only to one in this special case.)
Step 5. The elemental through variables are expressed in terms of the loop through variables as in
(12.11).
Step 6. The elemental across variables can be evaluated using the explicit constitutive relations given
in Table 12.1 and (12.9).
Note, however, that the loop formulation cannot be applied directly to models with sources of through
variable as the constitutive relation explicitly expressing their through variables necessary in Step 2 of
the formulation does not exist. Also, the integro-differential form of the equations (12.8) does not suit to
common computer computational routines for simultaneous solving of the equations.

12.3 Extended nodal formulation


12.3.1 Principles of the method
To overcome its limitations, the nodal formulation can be extended in such a way that the formulated
equations include not only the nodal across variables, but also the through variables of some of the
physical elements in the model. A moderate increase in the number of the resulting equations to be solved
simultaneously is outweighed by the applicability of the extended nodal method to models consisting of
any physical elements.
Let us first demonstrate that the equations for the system models in Table 12.1 resulting from the
extended nodal formulation can be expressed directly in the first-order algebro-differential form. This
form suits much better to computer computational routines than the integro-differential form of the
equations (12.8) derived using the plain nodal formulation.
The undesirable integration operation in (12.8) originated from substituting the constitutive relation
of the L element for its through variable i3 into (12.4) in Step 2 of the nodal formulation. We can avoid
this by leaving i3 in (12.4) without any change, and by adding the implicit constitutive relation of the L
element from Table 12.1 as a third equation. Then,
12.3. EXTENDED NODAL FORMULATION 12-5

d
J + C dt v2 − i3 = 0
i3 + Gv4 = 0 (12.13)
d
v3 − L dt i3 = 0

When writing the elemental across variables in terms of the nodal across variables (12.7) we obtain
the three first-order algebro-differential equations for three unknown variables vM , vN and i3

d
C dt vM − i3 = −J
i3 + GvN = 0 (12.14)
d
vN − vM − L dt i3 = 0

These equations can be expressed in the matrix form


     
Cs 0 −1 vM −J
 0 G 1  ·  vN  =  0  (12.15)
−1 1 −Ls i3 0

d
where the symbol dt is replaced by the operator s assuming that the product s x should be interpreted
d
as dt x.
The left-hand-side square and right-hand-side column matrices of (12.15) can be arranged in the
tabular form convenient for manual formulation of equations

vM vN i3
M Cs −1 −J
(12.16)
N G 1
L −1 1 −Ls
The empty entries in the tables stand for zeros.
In a similar manner, the nodal formulation can be extended towards sources of across variables or any
other physical elements. In the next paragraphs, you will learn how to form the extended nodal equations
very easily following a straightforward pattern.

12.3.2 Stamp submatrices of uncontrolled elements


The ‘stamp submatrices’ in Table 12.2 allow formulating the extended nodal equations as simply as
imprinting stamps. The submatrix entries should be added to the entries of the equation matrices in a
tabular form similar to that in the example (12.16). The element symbols in the table are mostly in their
generic multi-physical-domain form.
The Table 12.2 gives submatrices for formulation of the equations with elemental through variables
for all the physical elements listed in the table. It gives also submatrices for nodal equation formulation
without the elemental through variable for those elements for which such submatrices exist.
The initial conditions at the time instant t = 0+ shown in the right-hand-side submatrices of some
of the elements apply to the case when the resulting equations are considered as Laplace transforms of
algebro-differential equations. If this is the case, s represents the Laplace-transform operator.
The pattern of the extended nodal formulation exploiting the stamp submatrices is easily identified.
Step 1. Inspect the system model and count the non-reference nodes as well as the elements the through
variables of which should be introduced as additional variables in the formulated equations. These will
be

• sources of across variables

• inductors and springs

• sources of through variables, conductors, dampers, capacitors, and inertors if their through variable
is used to control an element in the model, or if it is requested as the output variable.
12-6 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

Table 12.2: Stamp submatrices of uncontrolled physical elements


Without With
Physical element
through variable through variable

a va ivb
Conductor va vb a 1
or damper a G −G
b −1
b −G G
b G −G −1

a va vb i
va vb
a 1
Resistor a 1/R −1/R
b −1
b −1/R 1/R
b 1 −1 −R
a va i vb
Capacitor va vb a 1
or inertor a Cs −Cs Cv(0)
b −Cs Cs −Cv(0) b −1
b Cs −Cs −1 Cv(0)
a va vb i
a 1
Inductor
or spring b −1
b 1 −1 −Ls −Li(0)

M i1 i2
Mutual L1 −M s −L1 i1 (0)
inductance −M i2 (0)
L L
−L2 i2 (0)
L2 −M s
−M i2 (0)
a i
Source of va vb
a 1
through a −J
variable b −1
b J
b 1 J
a va vb i
Source of
a 1
across
variable b −1
b 1 −1 E
12.3. EXTENDED NODAL FORMULATION 12-7

Table 12.3: Stamp submatrices of controlled elements


Without With
Physical element
through variable through variable

Source of through c ai vc vd i
vc vd
variable controlled by a 1
vc a g −g
across variable g b −1
b −g g
d b g −g −1

c ic a i ic i
Source of through ic
variable controlled by a 1
a β
through variable b −1
b b −β
d b β −1

c i a i va vb ic i
c
Source of across
variable controlled by a 1
through variable b −1
r
1 −1 −r
d b
c a i va vb vc vd i
Source of across
a 1
variable controlled by vc
across variable b −1
m
1 −1 −µ µ
d b
Ideal c va vb i
i
operational a 1
amplifier a b −1
d 1 −1
b

Step 2. If the number of he counted elements is m and the number of the non-reference nodes is n, draw
tables for the (m + n)-dimensional square and column matrices. Denote the first n columns of the square
table by the identifiers of the nodal across variables, and the remaining m columns by identifiers of the
computed through variables. At the same time, denote the first n rows of both tables by the identifiers of
the system model nodes and the remaining m rows by identifiers of the elements related to the computed
through variables.
Step 3. For each of the elements in the system model associate its poles a and b shown in Table 12.2
with the nodes of the system model. Add then the entries of the element submatrices to the related
entries of the resulting matrices. Ignore the entries related to the reference node.
For example, to form matrices (12.16) for the system models in Table 11.1, the poles of the involved
elements should be associated with the model nodes in the following way:
element J: a ↔ M, b↔0
element C: a ↔ M, b↔0
(12.17)
element L: a ↔ N, b↔M
element G: a ↔ N, b→0

Try now to form matrices (12.16) using the stamp submatrices of the elements.

12.3.3 Stamp matrices of controlled physical elements


The stamp-matrix approach can be easily applied also the models with controlled elements. If an element
is controlled by a through variable, this controlling through variable should be introduced as the additional
variable in the formulated equations.
12-8 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

Table 12.4: Stamp submatrices of ideal transducers


Without With
Transducer
through variable through variable

va vb vc vd i
i a 1
c a b −1
Ideal
c −n
transformer
d n
d b 1 −1 −n n

va vb vc vd i
va vb vc vd
i a 1
a g −g
c a b −1
b −g g
Ideal gyrator c g −g
c g −g
d −g g
d −g g
d b g −g −1

Table 12.3 gives stamp matrices of controlled sources. The controlling through variables are denoted
as iC , while iC stands for the controlling across variables between system model nodes c and d.

12.3.4 Stamp matrices of transducers


Table 12.4 demonstrates that the extended nodal method can be easily applied also to systems containing
such four-pole elements like an ideal transformer, an ideal gyrator, or an ideal operational amplifier.

12.3.5 Stamp matrices of blocks


The above given procedure for equation formulation can be applied also to system models represented
by block diagrams as well as by multipoles combined with blocks. Table 12.5 gives the stamp matrices
for basic blocks.

12.4 Block diagrams


12.4.1 Construction of block diagrams
Though block diagrams might look similar to multipole diagrams at the first sight, their principle is very
different. We shall demonstrate the procedure required for the block-diagram construction using again
the system examples given in Table 11.1.
Step 1: Formulation of a set of equations.
Any of the school-book pencil-and-paper procedures yielding consistent equations characterizing the
system model can be used for this purpose.
Let us utilize for our demonstration the implicit equations (12.3), (12.4) and (12.5) formulated in the
previous section.
Step 2: Decreasing the order of the differential equations.
Any n-th order differential equation can be converted to n first-order equations by substitution for the
higher-order derivatives.
There are no higher-order equations among those considered in our demonstration, however.
Step 3: Conversion of equations into explicit form.
Such a conversion should result in a set of equations each of which expresses explicitly one of the unknown
variables.
12.4. BLOCK DIAGRAMS 12-9

Table 12.5: Stamp submatrices of basic blocks


Block Stamp submatrix

a va
a 1
signal source

k va vb
b a a 1 −k
scalor

b va vb vc
a
c a 1 −1 1
comparator

b a va vb
a s −1 va (0)
integrator

b a va vb
a 1 −s −vb (0)
differentiator

b a vA vB x1 x2 x3
a 1 −b0 −b1 −b2 − sb3 −b3 x3 (0)
va b 3 s3 + b 2 s2 + b 1 s + b 0 1 −a0 −a1 −a2 − s −x3 (0)
= 3 s −1 x1 (0)
vb s + a2 s2 + a1 s + a0
s −a x2 (0)
transfer block

Table 12.1 shows the constitutive relations of the physical elements from our example in two different
explicit forms (except the through-variable source for which only one explicit form exists). So, we are
supposed to choose one of the explicit expressions for each of the elements and, at the same time, to
rearrange the relations (12.4) and (12.5) into the explicit form in such a way that there is just one relation
for each variable. Therefore, we are faced with the so called causality problem.
For computational reasons numerical integration is preferred to numerical differentiation in block-
diagram oriented packages. Let us just form the first three explicit relations as
i1 = J R
v2 = 1/C R i2 dt + v2 (t0 ) (12.18)
i3 = 1/L v3 dt + i3 (t0 )
Once we made this choice, it can be proven that the rest of the explicit relation must be formed in
the following way:
v4 = 1/G · i4
i2 = −i1 + i3
i4 = i3 (12.19)
v1 = v2
v3 = −v2 + v4
For example, if we had chosen the relation i4 = G v4 for the fourth element, we would not be able to
solve the problem. For some system configurations, even some of the integral relations must be replaced
by differential ones to solve the causality problem.
Step 5: Setting up the block diagram.
The resulting block diagram is shown in Fig. 12.2 using the DYNAST graphical notation. Each of the
blocks in the diagram represents just one of the explicit relations from (12.18) and (12.19) (no blocks are
needed for the trivial relations for i4 and v1 ).
12-10 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

i1
i2 v1 = v2
v3 i4 = i3
J
1/C v4
1/L
1/G

Figure 12.2: Block diagram representing the systems in Table 11.1.

12.4.2 Block diagrams vs. multipoles


Let us summarize the experience learnt from this simple example:

• If we want to represent energy interactions of a real subsystem by a block, for most of the subsys-
tems we must have at our disposal two different blocks at least, each of them representing the same
constitutive relation, but expressing explicitly different variables. In the case of the multipole ap-
proach, only one model is sufficient for each subsystem, and this model is reusable in any consistent
system configuration.

• In the block-diagram model of a complete system, in addition to the blocks representing the individ-
ual subsystems, there must be also blocks respecting the postulates of continuity and compatibility.
The latter blocks are not reusable as they suit just to one specific system configuration. For the
multipole models the postulate-respecting relations can be formed automatically and no additional
blocks are needed.

• To choose which block should model a particular subsystem, and to form blocks respecting the
postulates, we have to solve the causality problem. In the multipole approach such a problem is
not encountered at all.

The manual procedure for setting up a block diagram for the physical-level model of a real-life system,
usually nonlinear, is much more involved than could be shown in this example, of course. However, there
is also a big difference between DYNAST and the common block-diagram oriented packages in the way
of solving equations.
Once a block diagram is submitted to DYNAST, a set of implicit algebro-differential equations is
formed, and these are solved simultaneously as it is the case of the equations for multipole models. The
common block-diagram oriented programs usually do not solve any equations in reality. They would not
treat the relations (12.18) and (12.19) as equations, but as assignments.
The variables are processed successively by one block only at a time at discrete time steps. The
process starts from blocks with no input like the block specifying i1 (t1 ) ⇐ J(t1 ) in Fig. 12.2. But this
is not so simple in the case of feedback loops in the block diagram. For example, the summing block
for the output variable i2 (t1 ) ⇐ −i1 (t1 ) + i3 (t0 ) cannot specify the correct value of this variable at the
end of the first time step, but only after a number of time steps and iterations. This problem is even
more difficult if there is no integrator delaying the variable propagation in the feedback loop by one step.
Solving such algebraic loop problems requires special techniques.
In many cases, the blocks in a block diagram must be rearranged before the simulation starts as their
order is critical to the validity of the computed results. There are, however, system models characterized
by systems of differential equations the order of which is time-variable. Solving such an change-of-order
problem requires reconfiguration of the block diagram not only prior, but also during the simulation.
As all the equations are solved simultaneously in DYNAST, there are no algebraic-loop problem. As,
in addition, the equations are solved in DYNAST directly in their natural implicit algebro-differential
form, there is no change-of-order problem. Not only this, the related computational procedures for such
equations are in principle more efficient and robust then those implemented in the common block-diagram
packages. Obviously, developing software for automated construction of physical-level system models as
a pre-processor for a common block-diagram oriented package is not the optimal way of coping with the
problem.
12.5. LAGRANGE’S EQUATIONS OF MULTIPOLES 12-11

12.5 Lagrange’s equations of multipoles


Multipole models of subsystems can be characterized by constitutive relations formed using different ap-
proaches. Lagrange’s equations of second kind combined with undetermined multipliers can be exploited
to form the constitutive relations using the following procedure.
Step 1. Identify the subsystem energy entries and mass centers of the bodies within the subsystem.
Step 2. Determine the degree of freedom i of the subsystem bodies.
Step 3. Choose n ≥ i coordinates sj , j = 1, 2, . . . , n, to characterize motions of the entries and of the
body mass centers.
Step 4. Form r = n − i relations between the coordinates implied by the geometric constraints

fk (sj , t) = 0, k = 1, 2, . . . r (12.20)

Step 5. Differentiate (12.20) to obtain n − i relations for the subsystem geometric compatibility

d
fk (sj , t) = 0, k = 1, 2, . . . r (12.21)
dt

Step 6. Consider the ‘free’ diagram, i.e., consider the disconnected from a system. Show all the external
forces the system exerts on the to hold it in equilibrium.
Step 7. Formulate n Lagrange’s equations in the form
r
d ∂K ∂K ∂P ∂D X ∂fk
= + Qj − − − λk j = 1, 2, . . . , n (12.22)
dt ∂ ṡj ∂sj ∂sj ∂ ṡj ∂sj
k=1

where K, P, and D is the kinetic, potential, and dissipative energy of the. Qj is a generalized external
force associated with the generalized coordinate sj , whereas λk is the Lagrange’s multiplier representing
the generalized reaction force associated with the coordinate sk .
Step 8. By excluding the r multipliers λk from (12.22) obtain n − r equations for quasi-static balance
of inertial, potential and dissipative forces.

Example: Body with two revolute joints

Figure 12.3: (a) Body with two revolute joints, (b) its multipole model.

The subsystem shown in Fig. 12.3a is a body constrained to a planar motion with two revolute joints
denoted by A and B. O denotes the mass center of the body. Each of the joints is represented by two
poles in the multipole model of the body given in Fig. 12.3b. Poles Ax and Bx represent translational
motions of the joints along the x-axis whereas the poles Ay and By along the y-axis. Pole Oϕ corresponds
to the rotational motion of the body about the mass center O.
12-12 MODULE 12. FORMULATION OF SYSTEM EQUATIONS

The body shown in Fig. 12.3a is of i = 3 degrees of freedom. Let us consider n = 7 coordinates:
xA , yA , xB , yB , xO , yO , and ϕ. Then the r = n − i = 4 geometric constraints can be expressed as:
f1 = xO − xB − lB cos ϕ = 0 f2 = yO − yB − lB sin ϕ = 0
f3 = xA − xO − lA cos ϕ = 0 f4 = yA − yO − lA sin ϕ = 0
By differentiating the constrains we obtain the first part of the constitutive relations:
ẋO − ẋB + lB ϕ̇ sin ϕ = 0 ẏO − ẏB − lB ϕ̇ cos ϕ = 0
(12.23)
ẋA − ẋO + lA ϕ̇ sin ϕ = 0 ẏA − ẏO − lA ϕ̇ cos ϕ = 0
The body kinetic and potential energies are
1 1 1
K= mẋ2O + mẏO
2
+ JO ϕ̇2 P = mgyO
2 2 2
where JO is the moment of inertia of the body with respect to its center of mass.
Substitution into (12.22) yields for the individual coordinates:

xA : 0 = FAx − λ3 yA : 0 = FAy − λ4
xB : 0 = FBx + λ1 yB : 0 = FBy + λ2
d d
xO : dt mẋO = −λ1 + λ3 yO : dt mẏO = −mg − λ2 + λ4
d
ϕ: J
dt O ϕ̇ = τO − λ l
1 B sin ϕ − λ l
3 A sin ϕ + λ2 lB cos ϕ + λ4 lA cos ϕ
Finally, after excluding the multipliers, we are arriving at the equations for the force dynamic equi-
librium forming the second part of the constitutive relations for the multipole model of the body with
two revolute joints
mẍO = FAx + FBx
mÿO = FAy + FBy − mg (12.24)
JO ϕ̈ = τO + FBx lB sin ϕ − FBy lB cos ϕ − FAx lA sin ϕ + FAy lA cos ϕ

Example: Flexible crank-and-slide mechanism

Figure 12.4: (a) Slider-crank mechanism with a flexible connecting rod, (b) its multipole model.

Fig. 12.4a shows a slider-crank mechanism which is considered as massless, but its connecting rod is
assumed to be flexible with dissipative losses. In the corresponding multipole model given in Fig. 12.4b,
the rotational motion of the crank is represented by the pole Bϕ whereas the translational motion of the
slider represents the pole Ax .
The mechanism shown in Fig. 12.4a is of i = 2 degrees of freedom. Let us consider n = 4 coordinates:
xA , ϕA , ϕB and s. Then the r = n − i = 2 geometric constraints can be expressed as:
f1 = xA − lB cos ϕB + (lA + s) cos ϕA = 0
f2 = y0 + lB sin ϕB − (lA + s) sin ϕA = 0
By differentiating the constrains with respect to time we obtain:
ẋA + lB ϕ̇B sin ϕB + ṡ cos ϕA − (lA + s)ϕ̇A sin ϕA = 0
(12.25)
lB ϕ̇B cos ϕB − ṡ sin ϕA − (lA + s)ϕ̇A cos ϕA = 0
The potential and dissipative energies are in this case
1 2 1 2
P = ks D= bṡ
2 2
12.5. LAGRANGE’S EQUATIONS OF MULTIPOLES 12-13

where k and b is the stiffness and damping factor of the connecting rod, respectively.
Substitution into (12.22) yields for the individual coordinates:

xA : 0 = FAx + λ1
ϕB : 0 = τB + λ1 lB sin ϕB + λ2 lB cos ϕB
ϕA : 0 = −ks − bṡ + λ1 cos ϕA − λ2 sin ϕA
s: 0 = −λ1 (lA + s) sin ϕA − λ2 (lA + s) cos ϕA

Finally, after excluding the multipliers, we can complete the constitutive relations by the equations

τB − FAx lB (sin ϕB + tan ϕA cos ϕB ) = 0


−ks − bṡ − FAx (cos ϕA − tan ϕA sin ϕA ) = 0
12-14 MODULE 12. FORMULATION OF SYSTEM EQUATIONS
Module 13

Formulation of transfer functions

13.1 Formulation of system-response transforms


Let us consider a linear time-invariant dynamic system described in the Laplace transform domain by a
set of n algebraic equations. Such equations, resulting e.g. from the extended nodal formulation, are of
the general form
A(s) · X(s) = B(s) · U(s) + C(0+) (13.1)
where A(s) is a given n-dimensional square matrix, X(s) is the n-dimensional column vector of the system
variables to be determined, B(s) is the given n× p matrix, U(s) is the p-dimensional column vector which
represents the system excitation by independent sources, and C(0+) is the given n-dimensional column
vector representing the system initial conditions related to the energy accumulated in the system.
Obviously, the Laplace transform X(s) of the unknown vector x(t) can be written immediately in
terms of the inverse matrix A−1 (s) as
X(s) = A−1 (s) [B(s) · U(s) + C(0+)] (13.2)
After the inverse Laplace transformation is applied to (13.2), the vector of system responses x(t) can
be expressed as the sum of two response components
x(t) = xin (t) + x0 (t)
where
xin (t) = L−1 {A−1 (s) · B(s) · U(s)}
is the response of the system with zero initial conditions to an independent excitation, and
x0 (t) = L−1 {A−1 (s) · C(0+)}
is the response of the system with zero excitation from independent sources to system initial conditions.
Thus, the complete solution of (13.1) involves no more than the determination of the inverse of a
given matrix and the inverse Laplace transformation. Nevertheless, to calculate the inverse of a matrix,
say, anything more than a 3 × 3 matrix, is not a simple job without the help of a computer.
However, if we are required to find only one or two of the unknowns in (13.1), we need not compute
the inverse of the matrix A. Using the Cramer’s rule, the i-th component of X(s) in (13.1) can be
expressed as
∆i (s)
Xi (s) = (13.3)
∆(s)
assuming that the determinant ∆(s) of the matrix A(s) is nonzero. ∆i (s) is obtained form ∆(s) by
replacing the i-th column of ∆(s) by the right-hand-side of (13.1).

13.2 Determinant evaluation


Determinant ∆ of a 1 × 1 matrix A = [a] is defined as ∆ = a. Determinant of a 2 × 2 matrix
 
a11 a12
A=
a21 a22

13-1
13-2 MODULE 13. FORMULATION OF TRANSFER FUNCTIONS

is
a a
∆ = 11 12

= a11 a12 − a21 a22
a21 a22
Determinants of larger matrices can be computed by means of the Laplace expansion. In general,
expansion of the determinant ∆ of an n-dimensional square matrix A along its k-th row, resp. along its
l-th column, yields
n
X n
X
∆= akl (−1)k+l ∆kl , resp. ∆= akl (−1)k+l ∆kl (13.4)
l=1 k=1

where akl is the element in the k-th row and l-th column of the determinant, and ∆kl is the (n − 1)-st
order determinant obtained from the determinant ∆ by deleting the k-th row and l-th column of ∆.
For example, the determinant ∆ of the order n = 3 associated with the 3 × 3 matrix
 
a11 a12 a13
A =  a21 a22 a23 
a31 a32 a33

expanded along its first row



a a
∆ = a11 22 23 − a12 a21 a23 + a13 a21 a22


a32 a33 a31 a33 a31 a32

You may wish to show that expanding along another row or column gives the same result. This
method is valid for any n. Determinant evaluation using the Laplace expansion, however, is not suitable
for the analysis of large systems as it requires (n − 1)n! multiplications. While for n = 3 the number of
multiplications is 12, for n = 4 it becomes 72 already.
Fortunately, the system matrices are usually rather sparse so that the formulas (13.4) give us the
chance to take advantage of this matrix property. In general: the larger is the analyzed system, the higher
is the percentage of zero-value elements in its matrix A. By starting the determinant expansion from the
row or column with the least number of nonzero elements we can reduce the amount of computations
considerably.

13.3 Formulation of system transfer functions


In many design situations, rather than system responses, we need to compute a system transfer function
(or a matrix of such functions). Let us thus assume, that a system transfer function Hij (s) of a
system described by (13.1) is defined as
Xi (s)
Hij (s) = (13.5)
Uj (s)
where the dependent variable Xi (s), which is the i-th element of the vector X(s), is taken as the system
output, and the independent excitation variable Uj (s), being the j-th component of the vector U(s), is
considered as the system input. (According to the transfer function definition C(0+) = 0.)
Utilizing the Cramer’s rule (13.3), we can express the transfer function (13.5) as

∆i←j (s)
Hij = (13.6)
∆(s)

where ∆i←j (s) is the determinant of the matrix A(s) the i-th column of which has been replaced by the
j-th column of the v B(s).

13.4 Solved examples


Mechanical translation system. Consider the spring-mass-dashpot system shown in Fig. 13.1. Let
us obtain the transfer function of the system

y(s)
H(s) = (13.7)
F (s)
13.4. SOLVED EXAMPLES 13-3

Figure 13.1: Mechanical translational system.

by assuming that the force F (t) is the input and the displacement y(t) of the mass is the output.
The multipole model of the system is given in Fig. 13.1a. The node A represents the vertical motion of
the mass. The polarity of the input-force source has to comply with the chosen orientation of the velocity
ẏ(t). In this simple case only one equation is needed to describe the dynamics of the given system. This
is the equation in Fig. 13.1b are
ẏ F
A b + sm + k/s 1
The transfer-function H(s) can then be readily expressed as
1/s 1
H(s) = = 2 (13.8)
b + sm + k/s s m + sb + k

Mechanical rotational system. Consider the system shown in Fig. 13.2a. The system consists of a
load inertia and a viscous-friction damper driven by a rotational motor by means of a flexible shaft. The
motor is modelled as it is seen from the load.
System parameters:
τm = motor torque applied to the system
bm = motor damping (corresponding to the slope of its loading characteristics)
k = shaft torsional stiffness

Figure 13.2: Mechanical rotational system.

Jl = moment of inertia of the load


bl = load viscous-friction
ωl = load angular velocity
Let us obtain the transfer function of the system
ωl (s)
H(s) = (13.9)
τm (s)
The multipole diagram model of system is given in Fig. 13.2b. The nodes A and B represent the
rotational motions of the motor and the load, respectively. The equations describing the dynamics of the
system are
ωm ωl τm
A bm + k/s −k/s 1
B −k/s bl + sJ + k/s
13-4 MODULE 13. FORMULATION OF TRANSFER FUNCTIONS

The transfer-function H(s) can then be expressed as

∆21
H(s) = (13.10)

where
b + k/s −k/s
∆ = m

∆21 = −| − k/s| (13.11)
−k/s bl + sJ + k/s
Thus
k
H(s) = (13.12)
s2 Jbm + s(kJ + bm bl ) + k(bm + bl )

L-R-C circuit. Consider the electrical circuit shown in Fig. 13.3a. The circuit consists of an inductor
L, a resistor R, and capacitor C.

Figure 13.3: L-R-C circuit

If ei is assumed to be the input and v0 the output, then the transfer function of the system is assumed
to be
v0 (s)
H(s) = (13.13)
ei (s)
The multipole diagram of the system is given in Fig. 13.3b. It differs from the actual circuit diagram
by the voltage source representing the input excitation ei .
When formulating equations describing the system, we shall first take into consideration all the three
nodes shown in the diagram:

v1 v2 v3 i ei
1 1 −1
1
sL sL
2 −1 1
G+ −G
sL sL
3 −G G + sC
ei 1 1


1 −1

sL 0 1
sL

−1 1  
G+ −G 0 = G 1 + sC + C/L

∆(s) = sL sL



sL
0 −G G + sC 0


1 0 0 0

1 −1

sL 1
sL

3+1 −1 1
G
∆31 (s) = (−1) 0 = sL


sL G+
sL

0 −G 0
13.4. SOLVED EXAMPLES 13-5

The transfer function H(s) can then be expressed as

1
H(s) = (13.14)
LCs2 + RCs + 1
In this case we can considerably simplify the problem by formulating just two equations, one for the node
3 and the second one for the series combination of the elements ei − L − R. Then we shall obtain

v3 i ei
3 sC 1
ei − L − R 1 −sL − R 1

In this case
sC 1
∆(s) = = −(R + sL)sC − 1
1 −sL − R (13.15)
∆11 (s) = (−1)1+1 | − 1| = −1
The resulting transfer-function H(s) is identical to (13.14).

Seismograph. Fig. 13.4a shows a schema of seismograph. A seismograph indicates the displacement
of its case relative to the inertial reference. It consists of a mass m fixed to the seismograph case by a
spring of stiffness k, b is a linear damping between the mass and the case.
Let us define:
xi = displacement of the seismograph case relative to the inertial reference (the absolute frame)
x0 = displacement of the mass m within the seismograph relative to its case
y = x0 − xi = displacement of the mass m relative to the case.
Note that y is the variable we can actually measure.Since gravity produces a steady spring deflection,
we measure the displacement x0 of mass m from its static equilibrium position.
Considering xi as input an y as output, the transfer function is

Y (s) V0 (s) − Vi (s)


F (s) = = (13.16)
Xi (s) Vi (s)

k
y m

C M
M d
yM
vC m
k d
C
yC
(a) (b)

Figure 13.4: Seismograph

The multipole model of the seismograph is given in Fig. 13.4b. It is advantageous to take there as a
reference node the node C corresponding to the case instead of the node F corresponding to the actual
inertial reference. The transfer function can than be expressed as

vMC (s)
F (s) = (13.17)
vF C (s)

Exploiting the fact that the velocity between nodes F and C is constrained by the input velocity source
vi , we can write the matrix-form equations for the multipole diagram in the following way
13-6 MODULE 13. FORMULATION OF TRANSFER FUNCTIONS

vMC vF C vi
M b + sm + k/s −sm
F 0 1 −1

The determinants are in this case

∆(s) = b + ms + k/s ∆11 (s) = −sm (13.18)

From this
sm s2
F (s) = − =− 2 (13.19)
b + sm + k/s s + sb/m + k/m
Note that
lim F (jω) = 0 and lim F (jω) = −1 (13.20)
ω→0 ω→∞

Thus for low frequency inputs xi the mass m follows the case up and down. However, if the input
variable has a sufficiently high frequency, then the mass m remains almost fixed with respect to the
inertial reference and the motion of the case indicates the
p relative motion between the case and the mass.
For this reason the undamped natural frequency ω0 = k/m is made small in seismographs.

Liquid-level system. Consider the system of two interacting tanks shown in Fig. 13.5a.

Figure 13.5: Interacting tanks.

The fluid capacitance of the tanks is

A1 A2
C1 = and C2 = (13.21)
ρg ρg

where A1 , A2 [m2 ] are horizontal areas of the tanks, ρ [kg.m−3 ] is the mass density of the liquid, and
−2
g [m.s ] is the acceleration due to gravity. The volume fluid flow into the first tank through the control
valve is Qi . The tanks are interconnected by a short pipe with a valve representing a restriction of fluid
R1 . The fluid resistance of the second tank outlet is R2 .
The task is to compute the transfer function between the intake flow Qi and the outlet flow Q2 which
can be expressed as
Q2 p2
F (s) = = (13.22)
Qi R2 Qi
where p2 is the liquid pressure at the bottom of the second tank.
Based on the assumption that the system is either linear (for the laminar flow in the restriction) or
linearized (for the turbulent flow), it can be modelled by the multipole diagram given in Fig. 13.5 (the
fluid conductances G1 = 1/R1 , G2 = 1/R2 ).
The port diagram equations are then

p1 p2 Qi
1 G1 + sC1 −G1 1
2 −G1 G1 + G2 + sC2 0
13.4. SOLVED EXAMPLES 13-7

The relevant determinants are

∆(s) = G1 G2 + s[C1 (G1 + G2 ) + C2 G1 ] + s2 C1 C2 ∆21 (p) = G1 (13.23)

From there
∆21 1
F (s) = G2 = 2
(13.24)
∆ R1 C1 R2 C2 s + (R1 C1 + R2 C2 + R2 C1 ) + 1

x k1 u1
k1 k2
m1 k2 d
x1 1 2
f1 F1 F1
u2

m2 m1 m2
x2
(a) (b)

Figure 13.6: Multivariable mechanical system.

Multivariable mechanical system. Consider the system shown in Fig. 13.6a. We assume that the
system, which is initially at rest, has two inputs F1 (t) and F2 (t) and two outputs x1 (t) and x2 (t). In this
problem we would like to demonstrate formulation of a multivariable system transfer-function description
   
x1 (s) F1 (s)
= H(s) · (13.25)
x2 (s) F2 (s)

where the transfer-function matrix


 
H11 (s) H12 (s)
H(s) = (13.26)
H21 (p) H22 (s)

The equations describing the multipole diagram model of the system given in Fig. 13.6b are

ẋ1 ẋ2 F1 F2
1 b1 + sm1 + k1 /s −b1 −1
2 −b1 b1 + sm2 + k2 /s −1
The transfer-function matrix H(s) can then be expressed as
" #
1/s det A11 (s) det A12 (s)
H(s) = · (13.27)
det A(s) det A21 (s) det A22 (s)

where
det A = ∆(s) = (b1 + sm1 + k1 /s)(b1 + sm2 + k2 /s) − b21 (13.28)
and

det A11 (s) = −(−1)1+1 ∆11 (s) = |b1 + sm2 + k2 /s| = −b1 − sm2 − k2 /s
det A12 (s) = −(−1)1+2 ∆12 (s) = −| − b1 | = b1
(13.29)
det A21 (s) = −(−1)2+1 ∆21 (s) = −| − b1 | = b1
det A22 (s) = −(−1)2+2 ∆22 (s) = −|b1 + sm1 + k1 /s| = −b1 − sm1 − k1 /s

Clearly, the time responses x1 (t) and x2 (t) are given by

x1 (t) = L−1 [H11 (s) · F1 (s) + H12 (s) · F2 (s)]


x2 (t) = L−1 [H21 (s) · F1 (s) + H22 (s) · F2 (s)]
13-8 MODULE 13. FORMULATION OF TRANSFER FUNCTIONS

assuming zero initial conditions.


Show, that the equivalent of inertia and friction of the gear train referred to shaft 1 are given by

J1eq = J1 + n221 J2 + n221 n243 J3


b1eq = b1 + n221 b2 + n221 n243 b3

Similarly, the equivalent moment of inertia and friction of the gear train referred to the load shaft
are

J3eq = J3 + 1/n221 · J2 + 1/n221 · 1/n243 · J1


b3eq = b3 + 1/n221 · b2 + 1/n221 · 1/n243 · b1

The relation between J1eq and J3eq is thus

J1eq = n221 n243 J1eq (13.30)

and that between b1eq and b3eq is


b1eq = n221 n243 b1eq (13.31)
Note, that if n21 ≪ 1 and n43 ≪ 1 for a speed-reducing gear train, then the effect of J2 , J3 , b2 and
b3 referred to shaft 1 is negligible.

Rotary-to-linear motion conversion. In motion control problems, it is often necessary to convert


rotational motion into a translational one. For instance, a load may be controlled to move along a straight
line through a rotary motor and screw assembly, such as that shown in Fig. 13.7a.
Figure 13.7b shows a similar situation in which a rack and pinion is used as the mechanical linkage.
Another common system in motion control is the control of a mass through a pulley by a rotary prime
mover, such as that shown in Fig. 13.7c.

ωm
ωm M
M L
L

x x
(a) (b)

dm M L
L
M dl Fl
ωm
Jm
tau_m ml
x
(c) (d)

Figure 13.7: Rotary-to-linear motion conversion.

The system shown in Figs. 13.7a, 13.7b and 13.7c are all represented by a simple port diagram in
Fig. 13.7d. Elements τm , Jm and bm are modeling the drive motor there, and the elements Fℓ , mℓ and bℓ
model the mechanical load of the systems. The ideal transformer models the actual motion conversion.
13.4. SOLVED EXAMPLES 13-9

In case of the system shown in Fig. 13.7a, the transformer ratio n corresponds to the lead of the
screw, which is defined as the linear distance which the load travels per revolution of the screw. In the
other two systems n = r where r is the radius of either the pinion or of the pulley.
The equations describing the dynamics of the systems are following:
ωM vL F τm Fℓ
M b + sJm −n 1
L bℓ + smℓ 1 1
transformer n −1
The determinant of the system matrix is
∆ = bm + sJm + n2 (bℓ + smℓ ) (13.32)
Using this, we can easily derive the following transfer functions
vL n
=
τm bm + sJm + n2 (bℓ + smℓ )
ωM 1
=
τm bm + sJm + n2 (bℓ + smℓ )
F −n(bℓ + smℓ )
=
τm bm + sJm + n2 (bℓ + smℓ )
vℓ −n2
=
Fℓ bm + sJm + n2 (bℓ + smℓ )

From the second transfer function we can conclude that the equivalent torque τe , inertia Je and
damping be of the load that the motor sees is
τe = nFℓ , Je = n 2 m ℓ , and b e = n2 b ℓ
Thus any of the three rotary-to-linear conversion systems can be represented from the point of view
of loading the drive motor just by the elements τe , Je and be as shown in Fig. 13.7e.
Note also, that the third transfer function becomes
F 1
=−
τm n
for bm = Jm = 0 as could be expected.
From the last transfer function it is obvious, that the load sees the motor as shown in Fig. 13.7f where
the elements Fe , me and be are
Fe = τm /n, me = Jm /n2 , and be = bm /n2

Rotary-to-rotary motion conversion. Consider the system shown in Fig. 13.8a. In this system, a
load is driven by a motor through the gear train. Assuming that the stiffness of the shafts of the gear
train is infinite (there is neither backslash nor elastic deformation) and that the number of teeth on each
gear is proportional to the radius of the gear, obtain the equivalent moment of inertia and equivalent
friction referred to the motor shaft and referred to the load shaft.
In Fig. 13.8a, the numbers of teeth on gears 1, 2, 3 and 4 are N1 , N2 , N3 and N4 , respectively. The
angular displacements of shafts are ϕ1 , ϕ2 , ϕ3 and ϕ4 , respectively. Thus, ϕ2 /ϕ1 = −N1 /N2 = −n21 and
ϕ4 /ϕ3 = −N4 /N3 = −n43 . The moment of inertia and viscous friction of each gear train component are
denoted by J1 b1 , J2 b2 , J3 b3 , and J4 b4 , respectively. J1 b1 include the moment of inertia and friction of
the motor, J4 b4 include the moment of inertia and friction of the load.
The multipole model of the gear train system is given in Fig. 13.8b. The equations describing the
multipole diagram model of the system given in Fig. 13.8 are
ϕ̇1 ϕ̇2 ϕ̇3 τ2 τ4 τm τl
1 b1 + sJ1 n21 1
2 b2 + sJ2 1 n43
3 b3 + sJ3 1 1
T12 −n21 1
T34 −n43 1
13-10 MODULE 13. FORMULATION OF TRANSFER FUNCTIONS

Figure 13.8: Gear train system.

Você também pode gostar